This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
• bz/2mo, one attains m(i) ^ m o t 1 - *
(7)
That is ff = 1 - jfe/2. By use of the characteristic functions, the response propagator and correlation function have the generalized scaling forms which are valid for the times larger than t m ; c and for an arbitrary mo, Gp(t,f,mo)=p-2+"+^(pC(t),p$(t'),c(p-1,r'),¥'(p-1,mo))
Cp(tJ,m0)=p-2^g(pZ(t),pZ(t'),e(p-\T%
\x)f) and A(a?) are polynomials with the properties: deg.Q^~(a:) < (3M + l)L , deg.A(a;) < 2[f ], A(x) = A(-a;), A(0) = q21 + 1, and the following Bethe equation holds: i-l l
q- A(x)Q+(x)
L-l 1
= ] } ( 1 - MiQ-^QtixQ- ) 249
+ U.0- + xCj)Qt(xq)
.
(8)
2102
S.-S. Lin & S.-S.
Roan
Furthermore for 0
<
m
<
Qm(x),Qti-m(x)
M,
are
elements
in
By (iii) of the above theorem, the equation (8) for the sector m,N — m can be combined into a single one. For the rest of this report the letter m will always denote an integer between 0 and M: 0 < m < M. By introducing the polynomials Am{x),Q{x) via the relation, L-l
{Am{x), xm H(l-xNc?)Q(x))
= (q-mA(x),
(qmA(x),
Q+(x)),
Q+N_m{x))
,
3=0
the equation (8) for I = m,N — m becomes the following polynomial equation of Q(x),Am(x): L-l m
Am{x)Q{x) = q~
L-l 1
m
J J (1 - xcjq^Qixq- )
+q
J ] (1 + xCj)Q{xq)
,
(9)
with deg.Q(a;) < ML-m, deg.Am(ai) < 2[f-], Am(a;) = A m ( - a ; ) , A m (0) = qm + m q~ . The general mathematical problem will be to determine the solution space of the Bethe equation (9) for a given positive integer L. For L = 1,2, we have the following result. Theorem 2: (I) For L = 1, we have Am(x) = qm + q~m and the solutions Qm(x) of (9) form an one-dimensional vector space generated by the following polynomial of degree M — m, M m
~
Bm(x)
i
nm+i-l
_
g
„-m-i
= i + £ (IIq„m mi+ q-._m JL-i j=i
^+JW
i=i
(II) For L = 2, the equation (9) has a non-trivial solution Qm(x) if and only if deg.Qm(a;) = M — m + m' for 0 < m' < M. For each such m', the eigenvalue + g- m '- 2 )a; 2 CoCi + qm + q~ and Am(x) in (9) is equal to Am<m>(x) qi(q™'~i the corresponding solutions of Qm(x) form an one-dimensional space generated by a polynomial Bmtmt(x) of degree M — m + m' with -Bm>m-(0) = 1. For L = 3, this is the case related to the Hamiltonian (1). We consider the N x N matrix, / <5w-i U'N-I jV-2
w N-3
0
•••
u
S'iV-2 f
"iV-2
0
u V N-3
uS'N
N-3
0 0 \
U "JV-3
(10)
»i ^i u[ 0
\ o 250
w'0 v'0 6'0 j
Algebraic Geometry and Hofstadter Type Model 2103
with the entries defined by w'k — qk+ 2 +q k 2—qm—q m,v'k = (qk+*—q k ci+c 2 ), S'k = (qk~i +q~k~i){c0c1+c1c2 + c2Co), u'k = (qk-§ -q-k~§)c0c1c2. one can derive the following result.
2
)(co+ Then
Theorem 3: For L = 3, the condition of the eigenvalue Am(a;) = \mx2+qm+q~m, 0 < m < M, with a non-trivial solution Qm(x) in the equation (9) is determined by the solution of det(A — A m ) = 0, where A is the matrix defined by (10). For each such Am(x), there exists an unique (up to constants) non-trivial polynomial solution Qm{x) of (9) with the degree Qm(x) equal to 3M — m and Q m (0) ^ 0. 4. The Degeneracy and B e t h e Ansatz Relation of R o o t s of Bethe Polynomial We first discuss the degeneracy relation of eigenspaces of the transform matrix T(x) L
N
in <8> C * with respect to the Bethe solutions obtained in the previous section. As before, we denote A(x) the eigenvalues of T{x), whose constant term is given by L
T0 = D + l, where D := q~L
®Y®X) ®I®Z).
We have qD = {Z®X®I){X®I®Z){I®Z®X). We shall denote O3 the operator algebra generated by the tensors of X, Y, Z, I appeared in the above expression of T2. Then O3 commutes with D, hence one obtains a 03-representation on E 3 for each /. With the identification, U = D~^2Z® X® I, V = D'^^X®I®Z,0^is generated by U, V which satisfy the Weyl relation UV = uVU and the JV-th power identity. Hence O3 is the Heisenberg algebra and contains D as a central element. Then qD~T2 has the following expression, coCl{U + U'1) + c0c2(V + V-1) + Clc2(qD5/2UV
+ q^D-^V^U'1)
. 1
(11)
The above Hamiltonian is the same as HFK (1) with W — q~ D~ / V~ U~1, a = /3 = 7 = 1. Our conclusion on the sector m = M is equivalent to that in Ref. 6 as it becomes clearer later on. There is an unique (up to equivalence) non-trivial irreducible representation of O3, denoted by Cp , which is of dimension N. For each I, E3 is equivalent to AT-copies of C ^ as £>3-modules: E 3 ~ N C%. For 0 < m < M, we consider the space E 3 with ql = q±2m. The evaluation of E 3 on |x)±TO gives rise to a AT-dimensional kernel in E 3 . By Theorem 3, there are N polynomials Qm(x) of degree 3M - m as solutions of (9) with the corresponding N distinct eigenvalues Am(ar). The AT-dimensional vector space spanned by those Qm(x)s becomes a realization of the irreducible representation Cp for the Heisenberg algebra O3. 251
5 2
1
2104
S.-S. Lin & S.-S.
Roan
Now we discuss the relation between the Bethe equation (9) and the usual Bethe ansatz formulation in literature. For 0 < m < M, a solution Qm(x) in (9) always have the property Q m (0) ^ 0 by Theorem 3, hence one has the form a 3M-mQm(x) = U^=i~m(x ~ Tt) w i t h zi £ c * % setting x = zf1 in (9), we obtain the following relation among 2/s, which is called the Bethe ansatz relation, 2 < r +
,
3M-m
!TTiL±£L=
f^zi-Cj
TT
SfLli",
J^zi-qzn
t
< /<
3 M
-m .
For the sector m = M, the comparison of the ^-coefficients of (9) yields the expression of eigenvalue, 2M
AM = (q^~ +q~)s2
+ (qi - q~)si ^ z
n
+ (qi + q~ - q? -q~)^zizn
n=l
. Kn
With the substitution, p, = q^c^1^ = q^c^[l,p = q^c^1, the above expression coincides with (5.27) in Ref. 6. Note that the Bethe ansatz relation can be shown to be equivalent to the Bethe equation (9) for the sector M. However, the parallel statement is no longer true for other sectors m ^ M, i.e., it does exist some nonphysical Bethe ansatz solutions in the above form, while not corresponding to any polynomial solution of Bethe equation (9). Some example can be found in the (M - l)-sector. 5. High Genus Curves for the Hofstadter Model We are now going back to the general situation in Sect. 2. Note that the values ^ s of the curve C^ in (3) are determined by £Q and xN, denoted by y = xN, r] = $ . The variables (y, rj) defines the curve which is a double cover of y-line, % = Cn(yW where the functions A^,B^,CR,DR
+ (AR(y) - DR(y))V
- BR(y) = 0
are the following matrix elements,
(-At{v) B%(y) \ Jyf (-a? yb? \ Now we consider only the case: L = 3, ao = do = 0, bo = Co = 1, with generic hi, h2. The expression of T(x) is given by T(x) = x2(cia2X
®Z®Y
+ axb2Z ® Y ® X + bxd2Z ® X
equivalently, x~2D~T(x) is equal to the Hofstadter Hamiltonian (l) p =o with U = D~l/2Z
= 0,
Algebraic Geometry and Hofstadter Type Model 2105
which is an elliptic curve as a double-cover of the y-line. For the curve C^, the variables £o and £1 are related by ^ = £J~N, which implies that Cg can be identified with W x ZAT where W is a genus 6N3 — 6iV2 + 1 curve with the following equation in the variable p — (x, £o, £2)5 -tNaN .
w KV -
t-N ? 0
_
4- rNhN
S2 "l + x °1 _ JN ' TNj:NrN c X c; 2 l "l
-PNnN ^JV _ S2
4-
SO " 2 ^ „N cN„N x SO c 2
x
°2 JN ' 2
— a
By averaging the Baxter vectors \p,s) of C^ over an element p of VV, \p) := 1? S s =o IP> s)ls ' which defines the Baxter vector on W. Furthermore, the transfer matrix can be descended to one on W with the following relation, x-2T(x)\p) = |T_(p))A_(p) + |r + (p))A+(p) , where A± are the functions on W: A_(z, £0,6) = ( 8 6 C l - * j g o C a - < f a > , A+(x, &, 6 ) = g 2 ( ^ \ t a : - X t t - x t ) 2 C 2 ) - F o r a n eigenvector (
(£>£<)) £2)- Consider the -D-eigenspace decomposition of
2106 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13.
S.-S. Lin & S.-S. Roan M. Ya. Azbel, Sov. Phys. JETP 19, 634 (1964). R. J. Baxter, V. V. Bazhanov and J. H. H. Perk, Int. J. Mod. Phys. B4, 803 (1990). V. V. Bazhanov and Yu. G. Stroganov, J. Stat. Phys. 59, 799 (1990). W. G. Chambers, Phys. Rev. A140, 135 (1965). L. D. Faddeev and R. M. Kashaev, Comm. Math. Phys. 155, 181 (1995), hepth/9312133. P. G. Harper, Proc. Phys. Soc. London A 6 8 , 874 (1955). D. R. Hofstadter, Phys. Rev. B14, 2239 (1976). D. Langbein, Phys. Rev. 180 , 633 (1969). S. S. Lin and S. S. Roan, "Algebraic geometry approach to the Bethe equation for Hofstadter type models", cond-mat/9912473. R. Peierls, Z. Phys. 80, 763 (1933). P. B. Wiegmann and A. V. Zabrodin, Nud. Phys. B422, 495 (1994); JVucL Phys. B451, 699 (1995), cond-mat/9501129. J. Zak, Phys. Rev. A 1 3 4 , 1602 (1964); Phys. Rev. A134, 1607 (1964).
254
International Journal of Modern Physics B, Vol. 16, Nos. 14 & 15 (2002) 2107-2112 © World Scientific Publishing Company
LIMITATIONS O N Q U A N T U M C O N T R O L
ALLAN I. SOLOMON* Laboratoire de Physique Theorique des Liquides,
University
of Paris VI, France
SONIA G. SCHIRMERt Quantum Processes Group, Open University Milton Keynes MK7 6AA, United Kingdom
Received 17 October 2001 In this note we give an introduction to the topic of Quantum Control, explaining what its objectives are, and describing some of its limitations.
1. What is Quantum Control? The objectives of Quantum Control are • To determine quantum mechanical systems which will drive an initial given state to a pre-determined final state, the target state. • To describe—and hopefully implement—quantum systems which will through time evolution optimize given operator expectations corresponding to observables of the system. Among a wealth of applications are those to Quantum Computing, where it is clearly essential to be able to start off a quantum procedure with a given initial state, and to problems involving the population levels in atomic systems, such as the laser cooling of atomic or molecular systems. The mathematical tools necessary for the theoretical investigation of these control problems are diverse, involving algebraic, group theoretic and topological methods. The questions that one may ask include: (i) When is a given quantum system completely controllable? (ii) If a system is not completely controllable, how does this affect optimization of a given operator? (iii) How near can you get to a target state for a not completely controllable system? * Permanent address: Quantum Processes Group, Open University, Milton Keynes MK7 6AA, United Kingdom. Email: [email protected] tEmail: [email protected] 255
2108
A. I. Solomon
& S. G.
Schirmer
The answer to the first question depends on a knowledge of the Lie algebra generated by the system's quantum hamiltonian, that to the second arises from properties of the Lie group structure, while the last clearly involves ideas of topology. Especially in the area of Lie group theory, there is a large corpus of classical mathematics which can supply answers to questions arising in quantum control. In particular, for the type of controllability known as Pure State Controllability classical Lie Group theory has already given the basic results. The quantum control system we shall consider is typically of the form M
H = H0 + Y, fm(t)Hm,
(1)
m=l
where Ho is the internal Hamiltonian of the unperturbed system and Hm are interaction terms governing the interaction of the system with an external field. The dynamical evolution of the system is governed by the unitary evolution operator U(t,0), which satisfies the Schrodinger equation ih-U(t,0)=HU(t,0)
(2)
with initial condition {7(0,0) = I, where / is the identity operator. By use of the Magnus expansion, it can be shown that the solution U involves all the commutators of the Hm. The operators Hm, 0 < m < M, in (1) are Hermitian. Their skew-Hermitian counterparts iHm generate a Lie algebra L known as the dynamical Lie algebra of the control system which is always a subalgebra of u(N), or for trace-zero hamiltonians, su(N). The degree of controllability is determined by the dynamical Lie algebra generated by the control system hamiltonian H. If L = u(N) then all the unitary operators are generated and we call such a system Completely Controllable. A large variety of common quantum systems can be shown to be Completely Controllable. 1 ' 2 The interesting cases arise when L is a proper subalgebra of u(N). Such systems may still exhibit Pure State Controllability, in that starting with any initial pure state any target pure state may be obtained, as distinct from the Completely Controllable case, when all (kinematically admissible) states—pure or mixed—may be achieved. 2. Pure s t a t e controllability We shall restrict our attention here to finite-level quantum systems with JV discrete energy levels. The pure quantum states of the system are represented by normalized wavefunctions |\P), which form a Hilbert space H. However, the state of a quantum system need not be represented by a pure state |$) € H. For instance, we may consider a system consisting of a large number of identical, non-interacting particles, which can be in different internal quantum states, i.e., a certain fraction w\ of the particles may be in quantum state | * i ) , another fraction w^ may be in another state |*2) and so forth. Hence, the state of the system as a whole is described by a discrete ensemble of quantum states \^n) with non-negative weights wn that sum 256
Limitations
on Quantum
Control
2109
up to one. Such an ensemble of quantum states is called a mixed-state, and it can be represented by a density operator po on H with the spectral representation N
P0 = Y^Wn\^n)(^n\,
(3)
n=l
where {\$n) '• 1 < n < TV} is an orthonormal set of vectors in H that forms an ensemble of independent pure quantum states. The evolution of po is governed by p(t) = U(t,0)PoU(t,Q)\
(4)
with U(t, 0) as above. Clearly if all the unitary operators can be generated we have the optimal situation, complete controllability. However, classical Lie group theory tells us that even when we only obtain a subalgebra of u(N) we can obtain pure state controllability. The results arise from consideration 3 of the transitive action of Lie groups on the sphere Sk. The classical "orthogonal" groups 9(n, F) where the field F is either the reals SR, the complexes C or the quaternions %, are denned to be those that keep invariant the length of the vector v = {v\,V2, • • • ,vn); the squared length is given by v^v = £]™=1iTO, where Hi refers to the appropriate conjugation. These compact groups are, essentially, the only ones which give transitive actions on the appropriate spheres, as follows: (i) 9(71,3?) = 0(n) transitive on S^""1) (ii) Q(n,C) = U(n) transitive on S(2n~V (iii) ®{n,U) = Sp{n) transitive on Sf4™-1). Since we may regard our pure state as a normalized vector in CN and thus as a point on S^2N~l\ we obtain pure state controllability only for U(N) (or SU(N) if we are not too fussy about phases) and Sp(N/2), the latter for even N only. (Note that we cannot get 0(2N) as a subalgebra of U(N).) Complete controllability is clearly a stronger condition than pure state controllability. To illustrate our theme of the limitations on quantum control, we now give two examples based on a truncated oscillator with nearest-level interactions for which the algebras generated are so(N) and sp(N/2). Both these examples are generic.
3. Examples 3.1. Three-level oscillator
with dipole
interactions.
Consider a three-level system with energy levels E\, Ei, E3 and assume the interaction with an external field /1 is of dipole form with nearest neighbor interactions only. Then we have H = HQ + f(t)H\, where the matrix representations of HQ and 257
2110 A. I. Solomon & S. G. Schirmer
Hi are H0 =
Ei 0 0 " 0 E2 0 , 0 0 E3_
HX =
" 0 di 0 di 0 d2 . 0 d2 0
If the energy levels are equally spaced, i.e., E2 — E\ = E3 — E2 = /x and the transition dipole moments are equal, i.e., d\ = d2 = d then we have H'Q
-10 0 0 00 0 01
V
010 101 010
Hx
where H'Q is the traceless part of HQ. Both \H'0 and \H\ satisfy A + Af = 0
A J + JA = 0
where "0 0 1" 0 -10 .1 0 0.
J
which is a defining relation for so(3). The dynamical Lie algebra in this case is in fact so(3). It is easy to show that the matrix B = UAU^ is a real anti-symmetric representation of so(3) if U — U*J. Explicitly, a suitable unitary matrix is given by
u
1/V20
1/V2'
i/y/2 0
-i/y/2
0
i
0
Since the dynamical algebra and group in the basis determined by U consists of real matrices, real states can only be transformed to real states; this means that for any initial state there is a large class of unreachable states. This example is generic as it applies to iV-level systems, although for even N we need other than dipole interactions to generate so(N). The analogous dipole interaction generates sp(N/2) in the even N case, as we now illustrate.
3.2. Four-level
oscillator
with dipole
interactions.
Consider a four-level system with Hamiltonian H — HQ +
Ho
-Ei 0 0 0
0 0 -Eh 0 0 E2 0 0 258
0 0 0 Ex
f{t)Hi,
Limitations on Quantum Control 2111
and Orfi 0 0 di 0 d2 0 0 d 2 0 -di 0 0 -di 0 Note that \HQ and \H\ satisfy f x = - c,
xTJ + Jx = 0
(5)
for 0 0 0_-l
0 01' 0 10 100 0 0 0.
where J is unitarily equivalent to 0
IN/2
.—IN/2
0
(6)
which is a defining relation for sp(N/2). Consider an initial state of the form po = ix + OIN,
(7)
where x satisfies (5), it can only evolve into states pi = \y + aIN,
(8)
where y satisfies (5), under the action of a unitary evolution operator in exponential image of L. Hence, any target state that is not of the form (8) is not accessible from the initial state (7). Note that the initial state
Po
"0.35 0 0 0
0 0 0.30 0 0 0.20 0 0
0 " 0 0 0.15.
is of the form po = x + 0.25/4 and1 that that
IX = 1
0.10 0 0 0
0 0 0 0 0.05 0 0 --0.05 0 0 -0.10 0
satisfies (5). Consider the target state state
Pi
"0.30 0 0 0
0 0 0.35 0 0 0.20 0 0 259
0 " 0 0 0.15.
2112 A. I. Solomon & S. G. Schirmer
which is clearly kinematically admissible since
0 100] "0 1 0 0' 1000 1000 0 0 1 0 Po 0010 0001 0001. but note that pi=y
f
+ 0.25/4 and '0.05 0 0 0 " . _. 0 0.10 0 0 12/_1 0 0 -0.05 0 . 0 0 0 -0.10.
does not satisfy (5). Hence, p\ is not dynamically accessible from po for this system. Given a target state p\ that is not dynamically accessible from an initial state Po, we can easily construct observables whose kinematical upper bound for its expectation value can not be reached dynamically. Simply consider A = p\. The expectation value of A assumes its kinematical maximum only when the system is in state pi. Since p\ is not reachable, the kinematical upper bound is not dynamically realizable. 4. Conclusions In this short introduction to Quantum Control theory, we have described the goals of the subject briefly, and then illustrated the limitations by generic examples where complete control is not possible. The tools we have used are, in the main, those of classical Lie group theory. Theoretical problems that remain to be tackled include to what extent these non-controllable systems can, in fact, be controlled; and, of course, the paramount problem of implementing these controls in practice. Acknowledgements It gives us great pleasure to acknowledge many conversations with colleagues, including Drs. Hongchen Fu and Gabriel Turinici. A.I.S would also like to thank the Laboratoire de Physique Theorique des Liquides, of the University of Paris VI, for hospitality during the writing of this note. References 1. H. Fu, S. G. Schirmer, and A. I. Solomon. Complete controllability of finite-level quantum systems. J. Phys. A 34, 1679 (2001) 2. S. G. Schirmer, H. Fu, and A. I. Solomon. Complete controllability of quantum systems. Phys. Rev. A 63, 063410 (2001) 3. D. Montgomery and H. Samelson, Transformation groups on spheres. Ann. Math. 44(3), 454 (1943)
260
Internationa] Journal of Modern Physics B, Vol. 16, Nos. 14 & 15 (2002) 2113-2120 © World Scientific Publishing Company
RIGOROUS RESULTS O N T H E STRONGLY CORRELATED ELECTRON SYSTEMS B Y T H E SPIN-REFLECTION-POSITIVITY METHOD *
GUANG-SHAN TIAN Department of Physics, Peking University Beijing 100871, The People's Republic of China
Received 28 October 2001 In this talk, we shall briefly review some results on the strongly correlated electron systems, derived recently by applying Lieb's spin-reflection-positivity method. To explain the basic ideas of this method to a wide audience, we emphasize t h e important role played by Marshall's rule in studying t h e many-body systems Keywords: Strongly Correlated Electron Systems, Marshall's Rule, Spin-reflection-positivity Method
In the past several decades, the strongly correlated electron systems attract many physicists' interest. In particular, interplay between the itinerant magnetic orderings and the quantum transport properties of these systems is a main focus of the current research. To get insight into the strongly correlated electron materials, various models have been introduced and investigated. The best known examples are the Hubbard model, 1 the periodic Anderson model, 2 and the Kondo lattice model.3 As a concrete example, let us consider the Hubbard model. On a lattice A with N\ sites, the Hamiltonian of the Hubbard model has the following form
Jfff =
- * E E (^+dldi°)
+U
J2 (ftit -1) (% -\)-PN,
&
ieA ^
'
^
(i)
'
where c\a (cja) denotes the fermion creation (annihilation) operator which creates (annihilates) an electron of spin a at lattice site i. < ij > is a pair of lattice sites. t > 0 and U > 0 are parameters representing the hopping energy and the onsite Coulomb repulsion of electrons, respectively. In the following, we shall assume that, in terms of the Hamiltonian, lattice A is bipartite. In other words, it can be separated into two sublattices A and B, and electrons hops only from a site i in one sublattice to a site j in another sublattice. 'Dedicated to Professor F. Y. Wu's 70th birthday. 261
2114
G.-S.
Tian
These models enjoy some symmetries, which can be exploited to simplify their analysis. For instance, the Hubbard Hamiltonian HH commutes with the total particle number operator N = Xa(^it+"u)- Therefore, the Hilbert space of this model can be divided into numerous subspaces (V(N)}. Each of them is characterized by a conserved particle number JV. HH also commutes with the following spin operators
igA
ieA
i€A
2
Consequently, both S and Sz are good quantum numbers. Furthermore, when /x = 0, the ground state of the Hubbard Hamiltonian in the half-filled subspace with N = N\ is actually its global ground state. In this case, the Hubbard Hamiltonian has another symmetry: The pseudospin spin symmetry. More precisely, HH commutes with operators J+ = '%2 e (i)Cit g U' ^ " = Y. eCOcuCi-r, Jz = 2 5 Z ( « n + "14. - 1), i6A
igA
(3)
i6A
where e(i) = 1 for i £ A and e(i) = — 1 for i £ B. In literature, these operators are called the pseudospin operators. 4 In one dimension, the Hubbard model can be exactly solved by applying the Bethe ansatz. 5 However, in higher dimensions, this approach fails. Instead, various approximate analytical techniques were introduced and developed. Most of them are based on the mean-field theories.6 By applying effectively these methods, many interesting results on the strongly correlated electron systems have been derived. On the other hand, if it is possible, one would naturally like to re-establish some of these conclusions on a more rigorous basis. It will deepen our understanding on the electronic correlations in these models. In a seminal paper published in 1989, Lieb introduced a powerful method, the spin-reflection-positivity technique, to investigate the Hubbard Hamiltonian with an even number of electrons. 7 With this method, Lieb proved that the ground state of the Hubbard Hamiltonian at half-filling is nondegenerate and has total spin 5 = (1/2)\NA — NB\, where NA and NB are the numbers of the sites in sublattice A and B, respectively. Later, this technique was also applied to both the periodic Anderson model and the Kondo lattice model. 8 ' 9 Similar conclusions were established. Later, we applied this method to investigating the spin and the off-diagonal correlation functions, as well as the excitation gaps of these strongly correlated electron models. In a series of papers, we proved that • The on-site pairing correlation function of the negative-C/ Hubbard model is nonnegative. 10 More precisely, for any pair of lattice sites h and k, inequality
<*o(-C0 I 4 e h 4 ^ c k t I *o(-EO> > 0
(4)
holds true for the ground state of the negative-?/ Hubbard model in V(2M). This inequality confirms that the Bose-Einstein condensation in the negative-?/ 262
Rigorous Results on the Strongly Correlated Electron
Systems
2115
Hubbard model occurs at zero-momentum. • At half-filling, the spin correlations in the ground states of the periodic Anderson model and the Kondo lattice model are antiferromagnetic. 11 Very recently, by extending the spin-reflection-positivity method to the case of nonzero temperature, we also proved that, at any T ^ 0, the antiferromagnetic spin correlation in these models is dominant. 12 • On some special lattices subject to condition \NA — NB\ = 0(NA), such as the example shown in Fig. 1, the ground state of the Hubbard model has both the antiferromagnetic and ferromagnetic long-range orders. In other words, it is a ferrimagnet.13 • Define A q p = E0(NA + 1) + E0(NA - 1) - 2E0(NA) and As = E0(NA, S = 1) — EQ(NA, S = 0) to be the quasi-particle gap and the spin excitation gap of the Hubbard model, respectively. Then, relation A q p > Ag holds true. Similar inequalities were also proven for the periodic Anderson model and the Kondo lattice model.14 • Define A c = E0(J = 1)-E0(J = 0) = E0(NA+2)-E0(NA) to be the charged gap of these strongly correlated electron models, then inequality A c > A s holds. 15
Fig. 1.
The lattice structure of organic conjugated polymers.
Due to the page limit of this paper, it is impossible to discuss the spin-reflectionpositivity method and its applications in details. In the following, we shall briefly explain the basic ideas of it. As a matter of fact, this method is closely related to a very simple but important observation: After a proper unitary transformation, the ground state of a realistic quantum many-body Hamiltonian, in general, satisfies Marshall's rule,16 in a sense. To begin with, let us first consider a simple quantum mechanical system: One particle moving in a one-dimensional well, as shown in Fig. 2. The ground state ^o(x) of this system satisfies the Schrodinger equation
h2 d2ty0(x)
. .T , .
_
T
. ,
(5)
When v(x) — 0, $o{x) can be explicitly solved and we have ^o(x) = y/2/Rsm(Trx/R) > 0 for any 0 < x < R. This is Marshall's rule for <3>o in this special case. A direct corollary of this rule is that the ground state ^o(^) is nondegenerate. 263
2116
G.-S.
Tian
Fig. 2.
T h e potential function of a one-dimensional quantum well.
To show that ^?o(%) also satisfies the Marshall's rule when v(x) ^ 0, we notice that tyo(x) is a ground-state solution of Eq. (5), if and only if it is also a minimizing function of the following energy functional
n2 fRm{x)
™=LL
\
'*
dx
x)dx
(6)
Jo
in some function space -ff1(0, R), requiring that both l^l 2 and \dif)/dx\2 are integrable over (0, R). Assume that ^o(x) has indefinite sign in the interval, as shown in Fig. 3. Then, i
\ / /// // //
iV
k
l\
1/ \\
/1 //
\\ \\
A
/\ \\
i/
V=oo
« A /
1 \ l\ // / i l l // / / 11 1 1 1/
J—\—/-U-l / Fig. 3.
IW
V=o°
•
0
R
X
T h e ground state wave functions *o(^) and |*o|(z)-
we construct a new function |\Po|(2:) by taking the absolute value of $o(x). We notice that replacement of ^Q{X) with l^olOc) does not change the value of the second term in Eq. (6). On the other hand, it can be shown
/' Jo
dV0(x) dx
dx >
f
d | *o | (x) dx
dx
(7)
Jo (See the appendix of Ref. 17). Consequently, we have
min £(i/>) = £ ( * „ ) > £(\ *o I)
(8)
In other words, | \&o | (x) must be also a minimizing function of energy functional (6) and hence, a ground state solution of Eq. (5). 264
Rigorous Results on the Strongly Correlated Electron Systems
2117
Next, we show that the ground state wave function ^o{x) has no zero point in (0, R). If this is not true, then, there must be, at least, one point XQ € (0, R) such that |^o|(a;o) — 0. However, as a nonnegative solution of a second order elliptical differential equation, |\&o|(a;) satisfies the so-called Harnack theorem, which tells us that, on any open interval (a, b) C (0, R), there is a constant C(aj;,) such that max | *o I (&) < C ( o , b) min x£(a, b)
| * 0 I (x)
(9)
xE(a, b)
holds true. 18 Now, we take an open interval containing xo- Harnack theorem implies that l^oK^) = 0 in this interval. Repeating this argument an appropriate number of times, we find that | ^o | must be identically zero in (0, R). This is certainly absurd. Therefore, we reach the conclusion that ^o(x) = |*o|(a0 > 0 in (0, R) and is nondegenerate. The same ideas can be also applied to study the strongly correlated electron systems. First, let us consider a simple model: The antiferromagnetic Heisenberg model on a bipartite lattice. The Hamiltonian of this model is of the following form
HAF = £ JySi • 4 = 1 J2 J« ( ^ + 3 - + Si-Sj+) + Y, JaS-A <«>
(10)
with Jy > 0. Si is the spin operator at site i. The lattice A is assumed to be bipartite and connected by these coupling constants. Since Hamiltonian (10) commutes with Sz = ^2iS-lz, Sz is a good quantum number. A natural basis of vectors in subspace V(SZ = M) is given by Xa(M) = | m i , m 2 , •••, mNA)
(11)
where ra\ represents the eigenvalue of S-lz at site i and they are subject to the condition m\ + W2 H h TUNA = M. In terms of this basis, the ground state wave function ^o(M) of H\F can be written as *o(M) = ^ C Q X a ( M )
(12)
a
The sum is over all the possible configurations {xa{M)}. However, due to the positive sign of the coupling constants {Jy}, it is difficult to uncover directly the sign rule satisfied by {Ca}. To remedy this problem, one needs to introduce a unitary transformation U\ = exp [i^^Z^B S-lz 1, which rotates each spin in sublattice B by an angle 7r about its z-axis. 19 Under this transformation, HAF is mapped onto HAF = utHxeUi
- Yl ( - - V 2 ) (Si+Si-
+ 5,_5,+) + ] T J u S „ S J z
(13)
Notice that the coupling constants in the spin-flipping terms of HAF have negative signs. For HAF, we are able to show that the expansion coefficients {Ca} of its ground state ^o satisfy the Marshall's rule Ca > 0. First, by following the aforementioned 265
2118
G.-S.
Tian
steps, we re-write the ground state energy EQ as an expectation value of HAF in its ground state ^o a n d observe that, for any pair of indices a and a', (x«(M)
| ( s , + S j _ + 5 i _ 5 j + ) | Xa'{M))
>0
(14)
holds true. Consequently, the state |\l/o|, which is constructed by replacing each Ca with \Ca\ in Eq. (12), has a lower energy than \to- Therefore, it is also a ground state. Next, we consider the Schrodinger equation of |\Po|- By inequality (14), it can be shown that, if Ca = 0 for some index a, then any vector \a', which is related to Xa by a spin-flipping exchange, must have a zero coefficient in the expansion of l^o |- On the other hand, since the lattice is connected by the coupling constants {Jjj}, any vector X/3 c a n be reached from Xa by a finite number of spin-flipping exchanges. Therefore, by repeating the above process an appropriate number of times, we reach the conclusion that all the expansion coefficients are zero. That is impossible. As usual, the Marshall's rule satisfied by {Ca} implies that the ground state ^o(M) of i?AF is nondegenerate in subspace V(M). On the other hand, since ETAF is unitarily equivalent to the original antiferromagnetic Heisenberg Hamiltonian, we conclude that the ground state ^o(M) of the antiferromagnetic Heisenberg Hamiltonian HAF hi V(M) is also nondegenerate. Finally, we consider the Hubbard model. As explained above, we first introduce a unitary transformation which maps the original positive-?/ Hubbard model into a negative-?/ Hamiltonian. This can be achieved by the so-called partial particle-hole transformation U2,4 which is defined by ulctxUi = e{\)c\v
^ 2 t c u t> 2 = c u .
(15)
Under this transformation, the positive-E/ Hubbard Hamiltonian at half-filling (with Lb = 0) is mapped to
HH(-U) = -t £ £ (4,q* + i ^ ) - U £ (n,t - \) (nu - i) . (16)
ieA ^
'
^
'
When N, the number of electrons in the system, equals an even integer 2M, the ground state wave function ^o(2M) of HH(—U) can be written as
*o(2M) = X V
a
^®^
(17)
a, /3
where {i/^} are configurations of electrons of spin a defined by
V£ = < A r •••<,* I °>
(18)
In Eq. (18), (ii, 12, • • •, \M) denotes the lattice sites occupied by fermions of spin a. I 0) is the vacuum state. The total number of these configurations is Cff . In the Hubbard model, electrons are itinerant. Therefore, we have the so-called fermion sign problem. Consequently, the coefficients {WQ;g} do not satisfy the simple 266
Rigorous Results on the Strongly Correlated Electron Systems
2119
Marshall's rule. However, Lieb proved that, if one takes index a for row index and (3 for column index and writes the coefficients { W ^ } into a matrix W (with I r W ^ W = 1), then this matrix is, in fact, a positive-definite matrix. It is the generalized Marshall's rule for the ground states of the negative- U Hubbard model. To prove this fact, we follow the above procedure and rewrite EQ{—U) as the expectation of HH(-U) in ^0(2M). A little algebra yields Eo{-U) = 2Tr(TWW) - U ^
THV^MVWVi)
(19)
i€A
where T is the matrix of the hopping term of HH(—U) and N\ is the matrix of operator h\ — 1/2. Since matrix W is Hermitian, we can find a unitary matrix V, which diagonalizes it. Let {wm} be the eigenvalues of W and {| m)} be the column vectors of the diagonalizing matrix V. Then, EQ{—U) can be further reduced to E0(-U) = 2^(m
| f | m)w2m - U ^ Y ,
m
w
^n\{n \ N, \ m)\2
(20)
iGA m, n
Obviously, if we replace {wm} with their absolute values, the summations on the right hand side of Eq. (20) becomes less. In other words, by replacing the coefficient matrix W of * 0 (2M) with | W |, we obtain a new wave function | * 0 | ( 2 M ) , which has a lower energy than the ground state $o(2M). Therefore, it must be also a ground state of -fl#(—U). Furthermore, its coefficient matrix | W | is a semipositive definite matrix. By substituting |* 0 |(2M) into the Schrodinger equation HH(-U)\^0\ = Uol^ol and noticing that the lattice A is connected by electron hopping, we can further show that, if one eigenvalue wm = 0, then all the eigenvalues of W are equal to zero. This is impossible. Therefore, W is actually a positive definite matrix and the ground state $o(2M) is nondegenerate. Since HH(-U) is unitarily equivalent to HH{U) at half-filling, the ground state of the latter Hamiltonian is also nondegenerate. With the positive definiteness of the coefficient matrix W of $o(2M), inequality (4) can be proven as a direct corollary. Similarly, other results listed above are also based on this fact, although their proofs require a little more effort. In summary, the spin-reflection-positivity method reveals that the ground states of some strongly correlated electron systems satisfy the Marshall's rule in a more sophisticated manner. In return, this rule enables us to establish several important qualitative properties on the spin and superconducting correlations as well as the excitation gaps in these systems. Acknowledgement s I would like to thank Prof. Mo-Lin Ge for inviting me to the APCTP-Nankai Symposium. This work is partially supported by the Chinese National Science Foundation under grant No. 19874004. 267
2120
G.-S. Tian
References 1. J. Hubbard, Proc. Roy. Soc. London A 2 7 6 , 238 (1963); M. C. Gutzwiller, Phys. Rev. Lett. 10, 159 (1963); J. Kanamori, Pro. Theor. Phys. 30, 275 (1963). 2. For a review, see, for example, P. A. Lee, T. M. Rice, J. W. Serene, L. J. Sham, and J. W. Wilkins, Comments Condens. Matter Phys. 12, 99 (1986). 3. For a review, see, for example, G. Aeppli and Z. Fisk, Comments Condens. Matter Phys. 16, 155 (1992); H. Tsunetsugu, M. Sigrist, and K. Ueda, Rev. Mod. Phys. 69, 809 (1997). 4. C. N. Yang and S. C. Zhang, Mod. Phys. Lett. B4, 759 (1990). 5. E. H. Lieb and F. Y. Wu, Phys. Rev. Lett. 20, 1445 (1968). 6. For a detailed review on these methods, see E. Fradkin, Field Theories of Condensed Matter Systems (Addison-Wesley, Redwood City, California, 1991). 7. E. H. Lieb, Phys. Rev. Lett. 62, 1201 (1989). 8. K. Ueda, H. Tsunetsugu, and M. Sigrist, Phys. Rev. Lett. 68, 1030 (1992). 9. T. Yanagisawa and Y. Shimoi, Phys. Rev. Lett. 74, 4939 (1995); H. Tsunetsugu, Phys. Rev. B55, 3042 (1997). 10. G. S. Tian, Phys. Rev. B 4 5 , 3145 (1992). 11. G. S. Tian, Phys. Rev. B 5 0 , 6246 (1994). 12. G. S. Tian, Phys. Rev. B 6 3 , 224413 (2001). 13. S. Q. Shen, Z. M. Qiu and G. S. Tian, Phys. Rev. Lett. 72, 1280 (1994); G. S. Tian and T. H. Lin, Phys. Rev. B 5 3 , 8196 (1996). 14. G. S. Tian, Phys. Rev. B 5 8 , 7612 (1998). 15. G. S. Tian and L. H. Tang, Phys. Rev. B 6 0 , 11336 (1999); J. G. Wang and G. S. Tian, Commun. Theor. Phys. 34, 21 (2000). 16. W. Marshall, Proc. R. Soc. London A 2 3 2 , 48 (1955). 17. J. K. Freericks and E. H. Lieb, Phys. Rev. B 5 1 , 2818 (1995). 18. N. S. Trudinger, Comm. Pure Appl. Math. 20, 721 (1967). 19. E. Lieb and D. Mattis, J. Math. Phys. 3, 749 (1962).
268
International Journal of Modern Physics B, Vol. 16, Nos. 14 & 15 (2002) 2121-2127 © World Scientific Publishing Company
AN ALGEBRAIC APPROACH TO THE EIGENSTATES OF THE CALOGERO MODEL HIDEAKI UJINO Gunma National College of Technology, 580 Toriba-machi Maebashi-shi, Gunma-ken 371-8530, Japan
Received 9 November 2001 An algebraic treatment of t h e eigenstates of the (Ajv-i-)Calogero model is presented, which provides an algebraic construction of the nonsymmetric orthogonal eigenvectors, symmetrization, antisymmetrization and calculation of square norms in a unified way.
1. T h e Calogero Model In 1990's, one-dimensional quantum integrable systems with inverse-square longrange interactions 1-3 attracted renewed interests of mathematicians and physicists since their relationships with the theory of multivariable orthogonal polynomials 4 ' 5 are recognized. The (A/v-i-)Calogero model1 with distinguishable particles 6
where (Kjkf){-
• • ,xjt
• • • ,xk, • • •) = / ( • • • ,xk, •• -,Xj,- • •), j,k
G {1,2, • • •,JV}, is
known to have the nonsymmetric multivariable Hermite polynomial as the polynomial part of the joint eigenvector of the conserved operators.7~9 We introduce a transformed Hamiltonian whose eigenvectors are polynomials, %W := ( ^ ( a O r o
1
^ - EW) O
(2)
where the reference state and its eigenvalue are
4A\x) = n \x, -x fc rex P (-^ f; xi), l<j
EM =
m=l
^uN(Na+(l-a)).
We shall deal with the eigenvectors of the Hamiltonian (2) in C[x], the polynomial ring with N variables over C, while the eigenstates for the original Hamiltonian (1) 269
2122
H. Ujino
is in C[z]4 A ) := {f{x)<j>{^\x)\f G C[x]}. The Hamiltonian (2) is Hermitian with respect to the inner product on C[a;], oo
N
/
"[[Axi\4>(kA\x)\2f{x)g{x), f,geC[x], (3) where f(x) denotes the complex conjugate of f(x). The inner product is induced from the natural inner product on C[a;]0g '. The reference state corresponds to the weight function in the inner product (•, -}(A)The commutative conserved operators for the Hamiltonian are known to be the Cherednik operators. 10 To show this, we have to introduce the Dunkl operators, 11
and the creation-like and annihilation-like operators, M)\ ._ _ J_V{A) (A) _ J _ V ( A ) *' -Xl 2u,V< ' ai ~ 2 u l in G End(C[a:]), where the superscript t on any operator denotes its Hermitian conjugate with respect to the inner product (3). Prom these operators, a set of Hermitian and commutative differential operators, dj ' G End(C[x]), [dj ,d\. '] = 0, is constructed by N
k=j+l 10 12
which we call the Cherednik operators. '
The Hamiltonian (2) can be expressed
HM=UJ:(4A)-\a(N-l)). Thus we conclude that the Cherednik operators {cf- '\j = 1,2, • • •, A''} give a set of commutative conserved operators of the Calogero model. 2. The Nonsymmetric Multivariable Hermite Polynomial The Cherednik operators define inhomogeneous multivariable polynomials as their joint polynomial eigenvectors, which are nothing but the nonsymmetric multivariable Hermite polynomials that form an orthogonal basis of the polynomial ring C[:r]. 7 - 9 ' 13 Let / := {1, 2, • • • ,N - 1} and I := {l,2,---,iV} be sets of indices and let V be an iV-dimensional real vector space with an inner product (•,•). We take an orthonormal basis of V {ej\j G / } such that (£j,£fc) = Sjk- The A^-itype root system R associated with the simple Lie algebra of type AN~I is realized as R = {ej — Sk\j,k € I,j ^ k}(c V). A root basis of R is defined by 270
An Algebraic Approach to the Eigenstates
2123
II := {otj = £j — £j+i\j G / } , whose elements are called simple roots. Let R+ be positive roots relative to LT and R- :— — i?+. The root lattice Q is denned by Q := 0 g/Zccj and positive root lattice Q+ is defined by replacing Z with Z>oA reflection on V with respect to the hyperplane that is orthogonal to a root a is expressed by sa(p) := p — {a.v,p)a, where av := i^x is a coroot corresponding to a. The yl^-i-type Weyl group is generated by {SJ :— saj\aj G II} which is isomorphic to 6JV- For each w G W, we define Rw := R+ D w; -1 i?_. We denote by £(w) the length of w G W defined by £{w) := \Rw\. When w G W is expressed as a product of simple reflections, w = Sjk •••Sj2Sj1, with k — £(w), we call it reduced. By use of the reduced expression, the set R^, is given by R>l>
=
\ajl
' Sjl
\0lJ2 ) > ' ' ' J Sjl
S
J2
' ' ' Sjl
-1 \ajl
) J•
We introduce lattices P := © j e / Z > 0 £ j and P+ := {p = YLjzi Mj£j e -^Vi > M2 > • " > Miv > 0}, whose elements are called a composition and a partition, respectively. The lattice P is ^ - s t a b l e . The degree of the composition and partition is denoted by \p\ := X^e/£*j- Let W(p) := {w(/i)|io G W } be the VF-orbit of p G P . In a VF-orbit of W(/i), there exists a unique partition p+ G P+ such that p = w(p+) G P (W G W). We define p := § £ a £ f l + « = | E j e / ( ^ - 2j + l)£j and 1N := J2jei £r We introduce the following operator SA^X := J2jei ^i^j > ^ e ^*> j G /, which relates the Cherednik operators with lattice P , so that we can deal with the eigenvalues of the Cherednik operators in terms of the lattice P . We identify the elements of the lattice P with those of the polynomial ring over C, x* := x^x^2 •••x%N G C[x}. We denote the shortest element of W such that w~'i-{p) G P+ by Wy, and define p(p) := w^p). We introduce an order X on P ,
v
{y,pGP)^\v+<^
»?W(ii+) [ ( i - ^ e Q i Z/G W{p+),
d
where the symbol < denotes the dominance order among partitions, d
v
'
(ji,v G P+) <^>/i ^ A, \p\ = \v\ and J^ffe < k=i
'
^Pk, fe=i
for all I G / . The definition of the nonsymmetric multivariable Hermite polynomial is summarized as follows. Definition 1: The (monic) nonsymmetric multivariable Hermite polynomial hp ' G
SA»hM
= (A, p + ap{p) + L(N
271
- l)lN)hW.
(4)
2124
H. Ujino
Since SA^X is a Hermitian operator with respect to the inner product (3) and all the simultaneous eigenspaces of the Cherednik operators {d,(A^} are onedimensional in the sense that the eigenvalues of {d,(A^} are non-degenerate, it proves that the polynomials hfj, ' are orthogonal with respect to the inner product, i.e., (hfj. ,h{> )(A) = ^.I/H^M ||2- Actually, the nonsymmetric multivariable Hermite polynomials form an orthogonal basis in C[x\. 3. The Rodrigues Formula Here we show the Rodrigues formula that generates the monic nonsymmetric multivariable Hermite polynomial. We introduce the Knop-Sahi operators {e^A\ e^A^}7'14 and the braid operators {Sj '\j G 1} defined by c ^ := a{A^KlK2
• • • Ks-u
where Kj := Kjj+i,
j G / . The operators {SJ ', e^A^} intertwine the simultaneous
eigenspaces of {d^x}.
AU»
e™ = KN^
• • • K2Kxa^B)\
flf
> :=
[Khdf\
The raising operators {A^ ' \/j, G P+} are defined by
; = (S<^)S(J,B)
... s M,fl) e (A, B )ty >
j€L
Let SW/X be defined by Sw^ := 5,-, • • • Sj2Sj1, where w^ = Sjl • • • Sj2Sj1 is one of the reduced expressions of w^. Then we can show the following relations, = A^(SA»
d(mA^
[ 4 A ) t , 4 A ) t ] = 0,
+
for M,!/ G P + ,
A A
St£>d< > = d^J'-M (*)$<£, for /i G P,
(5)
which lead to the Rodrigues formula for the monic nonsymmetric Hermite polynomial. 9 Theorem 2: The monic nonsymmetric multivariable Hermite polynomial njj, with a general composition fi G P in the W-orbit of the partition /x+ G P+ is algebraically obtained by applying the raising operator A + and the product of braid operators SwJ to h^
— 1,
hiA) = {c^)^S^A^h[A\
h^
= ltfAS
w(/1+)>
„+
e P+>
where the coefficient of the top term is expressed as
TT l r r r - 7 -v^ v ^ ^ . = TT (« v ^ + + qp) $••= n ifH-«(« .p)). <£?== n
c ^)-=
Proof: Using Eq. (5), we can confirm that /i}j given by the above formula satisfies definition 1. • 272
An Algebraic Approach to the Eigenstates
Using the square norm for h0 ,, (A) , (A), ^ o ,fto ;CA) -
2125
— 1, (2?r)f (2w)iJV(JVo+(1_0))
j-rTjl+ja) JLl r ( i + a ) '
which is proved by a certain limit of the Selberg integral, 15 and the Rodrigues formula, we can calculate the square norm of the nonsymmetric multivariable Hermite polynomial in an algebraic fashion.9 Theorem 3: The square norms of the nonsymmetric multivariable Hermite polynomial hi ' with a general composition fj, 6 W(fj,+), fi+ € P+ is given by 9
=
^Wi^d-)HW j-r
11 (/3V ) M + + a^-a 2 ll r K + + a ^-^) + 1)
r ( ( a v , n+ + ap) + 1 + a ) r ( ( a v , /x+ + ap) + 1 - a)
Proof: The above formula can be verified by use of relations among raising and braid operators and evaluation of the eigenvalues of the Cherednik operators. • 4. Symmetrization and Antisymmetrization The nonsymmetric multivariable Hermite polynomials with compositions fi in the same W-orbit of the partition /j,+ share the same eigenvalue of the Hamiltonian (2), tfA)hW=u>\vL+\hW,
forMGW(M+),
/i+SP+.
More generally, the polynomial with compositions /j, € W(fi+) share the same eigenvalue of an arbitrary symmetric polynomial, e.g., any of the power sums, of the Cherednik operators. Thus any linear combinations of hjj, ', n G W(fi+), fi+ € P+ are joint eigenvectors of the Calogero Hamiltonian (2) and its higher-order conserved operators. Among all such linear combinations, we consider eigenvectors of the Calogero Hamiltonian (2) in VF-symmetric and PF-antisymmetric polynomial rings over C, C^] 1 * 1 ^. In our formulation, we do not use symmetrizer or anti-symmetrizer 7 which makes the coefficients of the top terms differ from unity. We introduce a sublattice of P+ by P+ +5 := {/J + <5|/Z e P+}, S := Y,J<EI(-^ ~J)£j t o describe the antisymmetric eigenvectors. Theorem 4: 9 Let # £ + ) + for /i+ € P+ and H^~ for n+ eP++6be the following linear combination of the nonsymmetric multivariable Hermite polynomials with compositions \i £ W(n+),
*#* = £ *£W, 273
(g)
2126
H. Ujino
whose coefficients are
Then the polynomials are in symmetric or antisymmetric polynomial rings over C, i.e., H^J G C[x]±w. We call them the symmetric and antisymmetric multivariable Hermite polynomials, respectively. Proof: Requiring the linear combination of the nonsymmetric polynomial (6) to satisfy KjHfr ' = ±HJi , b( + + = 1, we obtain the coefficients b + as shown in Eq.(7). D The symmetric and antisymmetric multivariable Hermite polynomials are identified by the polynomial parts of the eigenstates for all the conserved operators of the (j4iv_i-)Calogero model with indistinguishable (bosonic or fermionic) particles,1' 17 - 18 N
f)2
i
*
^M^EHS+^+JX: 3
J=l
_ O2' .^ t t
j,k=l
which is obtained by restricting the operand of the Hamiltonian (1) to the space C[*]±^)>^)|CM±lir^,=^)±(a). Prom the square norms of the nonsymmetric polynomials (/ij, %/iJ, )(A) and the coefficients bu+u , we can evaluate the square norms of the (anti-)symmetric multivariable Hermite polynomials. To prove the formula of the square norms, we need the following lemma, 16 Lemma 5: For p. G P+, we have an identity, v
(a ,p, H+ [IX t aap) f f i Ta fa
E/ N 1™ (a ,u + ap)±a 1 / v,, + \ + „. -pr
=
w
n
nn
(av,p + ap) *•*• (av,a + ap) ± a ' 11 j-r T-r
The lemma is proved by use of an expression of the Poincare polynomials.19 Theorem 6: 9 Let H{^)+ for p+ G P+ and H(^]~ for p+ G P+ + S. The square norms of the (anti-)symmetric multivariable Hermite polynomials are given by
T-r r ( ( a v , p + ap) + 1 T a ) r ( ( a v , / j + ap) ± a) J |
T((a\p + ap) + l)T((a^p + ap))
Proof: The proof is straightforward from theorems 3 and 4, and lemma 5. 274
•
An Algebraic Approach to the Eigenstates
2127
5. C o n c l u d i n g R e m a r k s We have presented the Rodrigues formula for t h e monic nonsymmetric multivariable Hermite polynomial which gives t h e n o n s y m m e t r i c orthogonal eigenfunctions of t h e ( J 4jv_i-)Calogero model with distinguishable particles. T h r o u g h (anti)symmetrization, we have constructed t h e (anti-)symmetric Hermite polynomials t h a t give t h e polynomial p a r t s of the eigenfunctions of t h e (^4jv-i-)Calogero model with distinguishable particles. T h e square n o r m s of t h e above three cases are calculated in a n algebraic manner. Our formulation in this work is also applicable t o the i?;v-Calogero models with distinguishable or indistinguishable particles a n d results in the Rodrigues formula for the monic n o n s y m m e t r i c multivariable Laguerre polynomial a n d square norms of the nonsymmetric a n d (anti-)symmetric multivariable Laguerre polynomials. 9 Acknowledgements I would like to express my sincere g r a t i t u d e t o Professor M. L. Ge for kindly inviting me to t h e symposium and to all t h e staffs a t Nankai University for their warm hospitality. I am grateful to Dr. A. Nishino for fruitful collaboration. I appreciate the Grant-in-Aid for Encouragement of Young Scientists (No. 12750250) presented by t h e J a p a n Society for the Promotion of Science for financial s u p p o r t . References 1. 2. 3. 4. 5. 6. 7. 8.
9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19.
F. Calogero, J. Math. Phys. 12 419 (1971). B. Sutherland, Phys. Rev. A 4 2019 (1971). M. A. Olshanetsky and A. M. Perelomov, Phys. Rep. 94 313 (1983). I. G. Macdonald, Symmetric Functions and Hall Polynomials 2nd. ed. (Oxford: Clarendon Press, 1995). C. F. Dunkl and Y. Xu, Orthogonal Polynomials of Several Variables (Cambridge: Cambridge University Press, 2001). A. P. Polychronakos, Phys. Rev. Lett. 69 703 (1992). T. H. Baker and P. J. Forrester, Nucl. Phys. B 492 682 (1997). H. Ujino and A. Nishino, Special Functions, Proc. Int. Workshop (Hong Kong, June, 1999) eds. C. F. Dunkl, M. Ismail and R. Wong, (Singapore: World Scientific, 2000) p. 394. A. Nishino and H. Ujino, J. Phys. A: Math. Gen. 34 4733 (2001). I. Cherednik, Inv. Math. 106 411 (1991). C. F. Dunkl, Thins. Amer. Math. Soc. 311 167 (1989). S. Kakei, J. Math. Phys. 39 4993 (1998). C. F. Dunkl, Commun. Math. Phys. 197 451 (1998). F. Knop and S. Sahi, Inv. Math. 128 9 (1997). M. L. Mehta, Random Matrices 2nd. ed. (San Diego: Academic, 1991). A. Nishino and M. Wadati, J. Phys. A: Math. Gen. 33 3795 (2000). H. Ujino and M. Wadati, J. Phys. Soc. Japan 6 5 2423 (1996). H. Ujino and M. Wadati, J. Phys. Soc. Japan 6 6 395 (1997). J. E. Humphreys, Reflection Groups and Coxeter Groups, (Cambridge: Cambridge University Press, 1990).
275
International Journal of Modern Physics B, Vol. 16, Nos. 14 & 15 (2002) 2129-2136 © World Scientific Publishing Company
ONSAGER'S ALGEBRA A N D PARTIALLY O R T H O G O N A L POLYNOMIALS
G. VON GEHLEN* Physikalisches Institut der Universitdt Bonn Nussallee 12, D-53115 Bonn, Germany
Received 20 January The energy eigenvalues of the superintegrable chiral Potts model are determined by t h e zeros of special polynomials which define finite representations of Onsager's algebra. T h e polynomials determining the low-sector eigenvalues have been given by Baxter in 1988. In the Z3—case they satisfy 4-term recursion relations and so cannot form orthogonal sequences. However, we show that they are closely related to Jacobi polynomials and satisfy a special "partial orthogonality" with respect to a Jacobi weight function. PACS: 02.30.1, 75.10.J, 68.35.R Keywords: Chiral Potts model, Integrable quantum chains, Onsager's algebra
1. Introduction F.Y.Wu and Y.K.Wang1 were the first to consider the Potts model with chiral interaction terms. Their interest in this generalization arose from duality considerations, but the idea proved to be very fruitful in many respects: Ostlund 2 and Huse 3 proposed the chiral Potts (CP) model for phenomenological applications: it allows to describe incommensurate phases using nearest neighbor interactions only. We give a few references4-7 from which the subsequent development can be traced, and turn directly to the superintegrable chiral ZN Potts quantum chain. 8 This is a particularly interesting model, because it provides some of the rare representations known for Onsager's algebra9 and in this sense generalizes the Ising quantum chain (for N = 2 it is the Ising model). Integrability by Onsager's algebra entails that all eigenvalues of the hamiltonian are determined by the zeros of certain polynomials, which for the chiral Potts model were first derived by Baxter. 10 Although the definition of Baxter's polynomials looks very simple, the properties of these polynomials turn out to be quite non-trivial and interesting. 11 The main part of this note deals with the properties of these polynomials. They satisfy N + 1-term recursion relations, therefore for N > 2 they cannot form orthogonal sequences. However, as found recently, 11 several properties which characterize orthogonal polynomials * e-mail: [email protected]. uni-bonn.de 277
2130
G. von Gehlen
are almost true for Baxter's polynomials (e.g. the zero separation property is true except for one extreme zero). We first recall the definitions of the superintegrable CP-hamiltonian and Onsager's algebra and then, following B.Davies, 12 we sketch how the formula for the energy eigenvalues emerges. We consider Baxter's polynomials and their recursion relations. Equivalent polynomials with their zeros in (—1,-1-1) for N = 3 are written in terms of a determinant. Their expansion in terms of Jacobi polynomials gives the surprising result that many of the expansion coefficients vanish, leading to the notion of "partial orthogonality". The hamiltonian defining the Zjv-superintegrable chiral Potts quantum chain 8 ' 8 is:
Here u — e 2 " ' ^ and Zj and Xj are Zjv-spin operators acting in the vector spaces C^ at the sites j = 1, 2 , . . . , L (L is the chain length). The operators obey ZiXj = Xj ZiUjSi<j; Zj* = X^ = 1 and we assume XL+I = X\ (periodic b.c). A convenient representation is (-Xj-)j,m = <Sz,m+i mod N and (Zj)iiTn — <Jjjmwm. For N > 3 the complex coefficients make the chain hamiltonian parity non-invariant. For TV = 2 we get the Ising quantum chain. For fixed JV there is only one parameter, the temperature variable k. Incommensurate phases arise due to ground state level crossings. H commutes with the Z^-charge Q = rjj=i Zj. We write the eigenvalues of Q as UJQ. Q = 0 , 1 , . . . , JV — 1 labels the charge sectors of %. We split H into two operators writing H^ = —^N(A0 + kAi). A remarkable property of H is that A0 and Ai satisfy 8 the Dolan-Grady 13 relations [i4o,[4>,[4),j4i]]] = 16 [AcAx];
[Au [Au [Au A0}}} = 16^,
A0],
which are the conditions 14 for A0 and A\ to generate Onsager's algebra A, which is formed9 from elements Am, Gi, m € Z, / G N, l> m, satisfying [Ai,Am] = 4Gi-m;
[GuAm]
= 2Am+i-2Am-i;
[GhGm]=0.
(2)
From (2), there is a set of commuting operators which includes %: Qm = | (Am + A-m + k(Am+1
+ A-m+i));
[Qi,Qm]
= 0;
Qo = 'H-
To obtain finite dimensional representations of A we require the Am (and analogously the G/) 1 2 ' 1 5 to satisfy a finite difference equation: ^2^-_n cik Ak-i = 0. This is solved introducing the polynomial (the main object of the present paper): n
T(z) = J2 a" zk+n k=—n 278
(3)
Onsager's Algebra and Partially Orthogonal Polynomials
2131
(from A the ak are either even or odd in k). Now the Am and Gm can be expressed in terms of the zeros Zj of T(z) and the set of operators Ef, Hf. A
m = 2 £ (
G
- = E (*™ - *7m) Hi-
(4)
Prom A these operators obey sZ(2, C)-commutation rules: [Ef, E^\
= Sjk Hk;
[H^ Ef] = ±2 Sjk
Ef.
So A is isomorphic to a subalgebra of the loop algebra of a sum of sl(2, C) algebras. Prom the first of eqs.(4) we can express "H in terms of the Zj and the operators Ef. Writing Ef = JXij ± iJyj, then in a representation Z(n,s) characterized by the polynomial zeros z\,... ,zn and a spin-s representation J^' of all the Jj, we get: (A0 + hA1)z{n,s)
= 2 X {(2 + k (zj + z?))j$.
= where J'xj
4 ^ 1
+
2 f c c
J +
+ i(Zj - zj1)
jW}
^ t !
is a rotated S'[/(2)-operator, and Cj = cosOj = \(ZJ + z~x).
For the CP-hamiltonians (1) the spin representation turns out to be s = \. Accordingly, all eigenvalues of (1) have the form # W = -N
( a + bk + 2 £™ = 1 mj
v/l+2fccos%
+ A;2) ,
m, = ±\.
(5)
a and b are non-zero if the trace of Ao and A\ is non-zero. 2. Baxter's polynomials No direct way is known to find the polynomials T(z) from the hamiltonian %. However, the invention of the two-dimensional integrable CP model, 16 ' 17 which contains W a s a special logarithmic derivative, and functional relations for its transfer matrix have enabled Baxter 10 to obtain the polynomials for the simplest sector of H, which at high-temperatures contains the ground state (the polynomials corresponding to all other sectors have been obtained subsequently in 1 8 - 2 0 ). Here we shall consider only the simplest case. Baxter 10 finds that in terms of the variable t or s = tN = (c — l ) / ( c + 1) (recall c = cos# of (5)) these polynomials take the form P
QL)(*) = j? E ( i T - Q
bJt)-**;
aQ,L = (N- 1)(L + Q) mod N. (6)
Here Q denotes the Zjy-charge sector. For Z3 eq.(6) is, written more explicitly: • i-\
P£\S)
-t~aQ,L
= !
_
{{f + t + 1)L + OjQ(? + UjH + W)h + U-Q(t2 + Ljt + U>2)L) . 279
2132
G. von Gehlen
Due to their Zjy-invariance t -> ut, the PQ' depend only on s = tN. The degree of the PQ '(S) in the variable s isfe^g= [((N — 1)L — Q)/N] where [a;] denotes the integer part of x. Considering sequences of these polynomials for fixed Q and L € N, we notice that the dimensionsft^Qdo not always increase by one when increasing L by one: at every Nth step the dimension stays the same: e.g. the dimensions of the PQ-0 for L mod N — 0 and L mod N = 1 coincide, see Table 1. The polynomials (6) have their zeros all on the negative real s-axis: % is hermitian and so in (5) we must have — 1 < Cj < +1 which means negative sj. We will prefer to deal mostly with equivalent polynomials in the variable c, defining
nW(c) = (c + i ) ^ p W ( s = f = i ) .
(?)
Our main concern in this paper is to learn about the properties of the IIQ ' (c) or Pk (S), e.g. whether these can be arranged into orthogonal sequences etc. We will find that the ITQ (C) are polynomials with quite remarkable properties. A number of special features of the UQ ' have been discussed recently. 11 Here we give some more detailed results for the Z3-case. As the recursion relations for the UQ ' contain a lot of information, we now show how to obtain these.
3. Recursion relations We start with the observation that the coefficients of the PQ \S) can be obtained from the expansion of (1 + 1 +12 +... + tN_1)L, simply by taking every N—th term of the expansion, starting with the coefficient of t(-N~1^L~Q. More precisely, we claim that we can define the PQ by the decomposition
^+t
+
e + ,,.+tN^)Lj}^\L V
^
N
^t.L,QpQL){s)
= J
(g)
Q=0
demanding the PQ to depend on s = tN only. Proof: Insert (6) into (8) to get: /
n
N \L l1 - tJ.N \
,.
w
. n-i.n—1 ,r1 ^ ^ u^+O)
^ = o j=o
(1 - uji t)L'
y
>
Interchanging the Q- and j-summations we see that the Q-summation gives zero for j ^ 0, leaving only the j = 0 term, which is (8). Eq. (8) can now be used to obtain recursion relations: Write N-l
(l + t + t2 + ... + tN~1)YJ
N-l
t°L'i PQL){s) = J2 t"L+^ PQL+1)(s).
Q=0
Q=0
280
(9)
Onsager's
Algebra and Partially
Orthogonal Polynomials
2133
Comparing powers of t, e.g. for L mod N = 0 this gives 1 1
/ 1 1 1
(i+i) :>(£+!)
S
I P^ \ (L) PI >w
(10)
s J \ p(L) j
V i
V^-W
\
For L mod N = k replace PQ —> PQ-k cyclically in both column vectors, keeping the same square matrix. Recursion relations not coupling Pk ' with different Q follow by the TV—fold application of these relations, leading to N + l-term recursion formulae. These can be transcribed into the corresponding formulae for the ITQ . For the Ising case Z2 these are of the Chebyshev type n(
L
+4)
4clXQ(i+2) + 4n!Q
0,
(11)
2 the ITQ form orthogonal sequences. However, for Z3 we have 11 and so for N (valid for all L > 0 and all Q):
n^ L+9) - 3(9c2 - 5)n^ L + 6 ) + 48 n^ L+3) - 64 n^L) = o
(12)
These n^ L) form 9 sequences, each labeled by (Lo, Q), where Q = 0,1,2 and L = 3J+LQ where LQ = 0,1,2, j = 0,1,2, The degrees of the polynomials appearing in this relation increase by two from the right to the left, but since the recursion is four-term, not three-term, these are not orthogonal sequences 21a . However, like (11) also (12) are of the most simple type b : all coefficients are independent of L. 4. Expansion in terms of Jacobi polynomials As the zeros of our IIQ are confined to and dense in the interval (—1,-1-1) we call this the basic interval like for orthogonal polynomials. Trying to determine (numerically) a weight function by the ansatz /_ 1 (1 + c ) a ( l — c)^ckU.Q (c) = 0 for k < bQtL fitting a and (3, we find (in the following we concentrate on the Z3-case) that there is an approximate solution very close to a = —/3 = | , but a and (3 come out to be slightly L and fc-dependent, in contrast to what is needed for orthogonality. However, for L —> oo and small fixed k, the solutions converge 1
(-
(L)
—-)
towards a = —f3 = 3 . So the n ^ seem to be close to Jacobi polynomials P^ 3 ' , but can we formulate an exact relation valid for finite L? Is there an exact property of the U.Q ' which replaces orthogonality? Numerical calculations 11 gave the surprising result that seemingly complicated HQ ' a If we consider Q = L mod 3, then we have only polynomials in c 2 , and the degrees 6jr,,Q are consecutive in powers of z = c 2 ("simple sets of polynomials") with integer coefficients. b For higher N we get similar N + l-term relations, e.g. for Z4: 4) +12) - 128(14c 2 - 1 7 ) n ^ + 8 ) - 2 0 4 8 c I I ^ + 4 ) + 4096 n ^ = 0. n (L+i6) _ 4 { 6 4 c 3 _ 5 6 c ) n < n ^
281
2134
G. von
Gehlen
Table 1.
Examples of Z3—polynomials rig '(c) and their Jacobi-components. Z3,
Q = L + 1 mod 3,
(a,/3) = ( § , - 1 )
n
L
2-["/3in^(c)
3
9c + 3
[0,|]
4
27c2-18c-5
5
81 c 3 + 2 7 c 2 - 5 7 c - l l
[3,-f,fl [o--f-o,iJ
6
35c3+81c2-135c-21
[0,0,0,^]
7
3 6 c 4 - 2 - 3 5 c 3 - 540c 2 + 2 7 0 c + 43
fo l°!
8
3 7 c 5 + 3 6 c 4 - 2754 c 3 - 702 c 2 + 711 c + 85
9
3 8 c 5 + 3 7 c 4 - 7290 c 3 - 1782 c 2 + 1593 c + 171
[0,0,0,-2,0,^] [0u u0 0 - 5 1 0u 22^1
10
3 9 c 6 - 2 • 3 8 c 5 - 25515 c 4 + 14580 c 3
2 7 171 4 ' 14 '
729 7291 40 ' 70 J
L i i "i 40 1 > 56 J ro 2 7 135 243 9477 L°> 4 i 14 > 20 ' 385 ' 37 38 i 56 ' 3081
+7965 c 2 - 3 1 8 6 c - 3 4 1
can be written as a combination of just very few Jacobi polynomials, e.g. n(!12) = 3 n c 7 + 3 1 0 c 6 - 5 - 3 1 0 c 5 - 80919c4 +140697c 3 + 27459c2 -16839c - 1 3 6 5 = — ( l Pft~*] 728
. For polynomials TT(C) of degree n we use c :
- P^~^\ 7T(C) =
[TTo, TTi, . . . , 7Tn] = ^ 7 r
f e
P^.-^)
( c )
(13)
fc=0
and define a scalar product with Baxter's variable t = ((1 — c)/(l + c)) 1 / 3 (see (6)) as the weight function (here we will not need to specify the normalization):
WW -£'*(££
1/3 v r(l)r r((2~'(cj )f 7r ^(c)7T
0
- /— oo
6s Vx
;irW(c(s))irW(c(s)).
The second part of this definition shows that it makes sense also if we prefer to use polynomials in the variable s, and that it preserves the original Z3—symmetry. Since the Ilk ' satisfy the recursion relations (12) they can be written as determinants of band matrices with a bottom line specifying the initial conditions (which are the lowest L polynomials). Omitting the bottom line, we define j x j—band matrices and their determinants Rj = det | band( [64, 48, 3p, 1, 0], j)\ , e.g.
Re
3p 1 0 0 0 0 48 3p 1 0 0 0 64 48 3p 1 0 0 0 64 48 3p 1 0 0 0 64 48 3p 1 0 0 0 64 48 3p
where p = 9c — 5, so that the polynomials Rj depend only on c . Now we get the nine sequences of the IIQ ' as linear combinations of the Rj which satisfy appropriate c
For Jacobi series as a generalization of Taylor series, see Ch.7 of Carlson. 2 3 We use the standard normalization of Jacobi polynomials, see e.g. Rainville. 22 282
Onsager's Algebra and Partially Orthogonal Polynomials 2135 initial conditions. Abbreviating Qj = Rj + 8Rj-i l4 3,-) = 5 (Qj ~ 8 < 2 J - I + 16Qj-2);
these are found to be: n ( 3 j ) = 9 c ^ _ i ± 3Q,-_i; 2
n |3i+l) = ± 1 8 c ^ . _ 1 +
Q^. +
2 Q j l
.
n (3j+l) =
n l 3 j + 2 ) = 3c(Rj - 4R,-_i) ± (Q,- - 4Q j _i);
^. _
A Q j i
.
(M)
n^ 3 j + 2 ) = 3Q,-.
For j — 1,2 use i?o = 1; -ft-i = R-2 = 0. From (14) we see that only two of the 9 sequences are independent. There are relations like e.g. IJ^j+1'—U.i:'+1' = 2E4 . To get the Jacobi-expansion of the Ilk , we only need to expand Rj and cRj. Using Jacobi-components denned in (13), from explicit calculation (for j < 36), we find that for k < j (only there) the j-dependence of the (Rj)k obeys the simple rule: (Rj)k
= (-3) fc k\ (2k + 1) {(-8)1 ak + 4? rk} ;
1
2(-3)m
_
3llJUo(3n + l)
_
_
3 n ' = m ( 2 n + l)
2(-3)m~2 n*=m(2n)
It follows that the ak do not contribute to the A; < j-components of Qj, and, using the recursion relations for the Pk' \(Qj)k
= -(cRj)k
2
= \(c Rj+1)k
3
(c), we conclude that for k < j we have
= V(-3)k +
k\ (2k + l)rk.
(ny+\ = (n?\ = (ui* \ = (n^) i
It further follows that +2
=(n^ >) =o for k<j. k
i
k
All k < j components of II^3 , II2 3 and U^^2' are proportional to (Qj)kWe get zero overlap between a polynomial YLQ ' of degree bi,tQ with all polynomials ITQ, which have at least b/^Q vanishing low-fc components (these can only be polynomials which have &£/,Q' > 2bL:Q). One of many such relations is e.g. ( n g j ) I n g / ' } ) = 0 for Q = Q' = 1
and 2j < f - 1.
This property may be called "partial orthogonality". Further rules, this time valid for all k, regard the vanishing of many components for particular linear combinations: Defining Qf
= n<3j'> =9cRi-1+3Qj-1
Q{1] =
Qj +9cRj-!
- Qj-i
= U+\
a<+\ ..., a g ^ ] ;
= [oi-\
a{-\
...,
a{-] ],
we have checked up to j = 30 that ak ' = 0 for k < j and that all even (odd) k Jacobi components of Q+ (Q_ ) vanish. So we conjecture for all j , j '
(Q{i)\Q{P) = 0. One can check some special cases of these results in Table 1. It has been found numerically11 that a similar partial orthogonality appears also for the Zjv-Baxter polynomials with N = 4, 5, 6. For even values of N further relations emerge, but these will not be discussed here. 283
2136
G. von Gehlen
5. C o n c l u s i o n The polynomials which play a central role for t h e calculation of the energy eigenvalues of the superintegrable Z3—chiral P o t t s model are found to be related to Jacobi polynomials in a very peculiar way. M a n y integrals giving the Jacobi-coefficients of Baxter's polynomials a r e found t o vanish. T h e s e observations should have a deeper group-theoretical b a c k g r o u n d , but t h e underlying symmetry is not yet clear to us. By reducing the formulation of the problem t o some basic facts, the present analysis tries to prepare t h e g r o u n d for clarifying t h e s y m m e t r y involved.
Acknowledgements The author is grateful t o Prof. Mo-Lin Ge for t h e very friendly hospitality during the A P C T P 2001 N a n k a i Symposium in Tianjin. T h i s work has been supported by INTAS-OPEN-00-00055.
References 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23.
F.Y. Wu and Y.K. Wang, J. Math. Phys. 17, 439 (1976) S. Ostlund, Phys. Rev. B24, 398 (1981) D.A. Huse, Phys. Rev. B24, 5180 (1981) D A . Huse and M.E. Fisher, Phys. Rev. B 2 9 , 239 (1984) P. Centen, M. Marcu and V. Rittenberg, Nucl. Phys. B205, 585 (1982) S. Howes, L.P. Kadanoff, and M. den Nijs, Nucl. Phys. B215 [FS7], 169 (1983) H. Au-Yang and J.H.H. Perk, Int. J. of Mod. Phys. B l l , 11 (1997) G. von Gehlen and V. Rittenberg, Nucl. Phys. B 2 5 7 [FS14], 351 (1985) L. Onsager, Phys. Rev. 65, 117 (1944) R.J. Baxter, Phys. Lett. 133A, 185 (1988). G. von Gehlen and S.-S. Roan, in "Integrable structures of exactly solvable twodimensional models etc.", NATO Science Series 7735, 155, Kluwer Acad. Publ.(2001) B. Davies, J. Phys. A: Math. Gen. 23, 2245 (1990) L. Dolan and G. Grady, Phys. Rev. D 2 5 , 1587 (1982) J.H.H. Perk, in Theta Functions Bowdoin 1987, Am. Math. Soc. (1989) E. Date and S.-S. Roan, J. Phys. A: Math. Gen. 33, 3275 (2000) H. Au-Yang, B.M. McCoy, J.H.H. Perk, S. Tang and M.L. Yan, Phys. Lett. 123A, 219 (1987) R.J. Baxter, J.H.H. Perk and H. Au-Yang, Phys. Lett. 128A, 138 (1988) G. Albertini, B.M. McCoy and J.H.H. Perk, (1989) Adv. Stud. Pure Math. 19, 1 (1989) S. Dasmahapatra, R. Kedem and B.M. McCoy, Nucl. Phys. B257 506 (1993) R.J. Baxter, J. Phys. A: Math. Gen. 27 1837 (1994) T.S. Chihara, An Introduction to Orthogonal Polynomials, Gordon and Breach (1978) E.D. Rainville, Special Functions, Chelsea Publ. Comp. (1971) B.C. Carlson, Special Functions of Applied Mathematics, Academic Press (1977)
284
International Journal of Modern Physics B, Vol. 16, Nos. 14 & 15 (2002) 2137-2143 © World Scientific Publishing Company
T H E R M O D Y N A M I C S OF T W O C O M P O N E N T B O S O N S IN O N E D I M E N S I O N
SHI-JIAN GU Zhejiang Institute
of Modern Physics, Zhejiang University,
Zhejiang Institute
of Modern Physics, Zhejiang University,
Zhejiang Institute
of Modern Physics, Zhejiang University,
Zhejiang Institute
of Modern Physics, Zhejiang
Hangzhou
310027, P.R.
China
Hangzhou
310027, P.R.
China
Hangzhou
310027, P.R.
China
Hangzhou
310027, P.R.
China
YOU-QUAN LI
ZU-JIAN YING
XUE-AN ZHAO University,
Received 3 December 2001
On the basis of Bethe ansatz solution of two-component bosons with SU(2) symmetry and 5-function interaction in one dimension, we study t h e thermodynamics of the system at finite temperature by using the strategy of thermodynamic Bethe ansatz (TBA). It is shown that the ground state is an isospin "ferromagnetic" state by the method of TBA, and at high temperature the magnetic property is dominated by Curie's law. We obtain the exact result of specific heat and entropy in strong coupling limit which scales like T at low temperature. While in weak coupling limit, it is found there is still no Bose-Einstein Condensation (BEC) in such I D system.
1. Introduction A two-component Bose gas has been produced in magnetically trapped 87Rb by rotating the two hypernne states into each other with the help of slightly detuned Rabi oscillation field.1 It was noticed2 that the properties of such Bose system can be different from the traditional scalar Bose system once it acquires internal degree of freedom. Bethe ansatz solution of SU(2) two-component bosons in one dimension was obtained. 3 ' 4 It was pointed out that the ground state of such a system is an isospin "ferromagnetic" state 4 which differs from that of spin-1/2 fermions in one dimension, such as SU(2) Hubbard model, 5 etc. An interacting SU(2) boson field trapped in a one dimensional ring of length L 285
2138 S.-J. Gu et al. can be modeled by the following Hamiltonian H=
dx
(i)
/ •
a
a,b
where a,b = 1,2 denotes the ^-component of isospin. The Bethe ansatz equations (BAE) of eq. (1) are obtained as follows ikjL _ _ TT *j — h+ic -I-T kj — \y — ic/2 e ~ M kJ -h-ic 1 1fc,-- Xv + ic/2
l=\ ^~ki+
ic/2 1=1 A7 - \„ - ic
U
where M denotes the total number of down isospins. Eq. (2) differs from the BAE of scalar bosons with periodic condition. 6 The second equation of eqs. (2) of isospin rapidity A arises from the application of quantum inverse method, 7 which can be inferred from spin-1/2 fermions8 too. However, the symmetry of bosonic wave function gives the first term on the right side of first equation of eqs. (2), which does not appear in the BAE for fermions. 2. Thermodynamics at Finite Temperature The strategy we use here is the thermodynamic Bethe ansatz (TBA) which was pioneered by C. N. Yang and C. P. Yang for the case of the delta-function Bose gas. 1 0 It is used to derive a set of nonlinear integral equations called TBA equations, which describe the thermodynamics of the model at finite temperature. Moreover, the As can be complex roots which should form a "bound state" with other As11 when T ^ O , which arises from the consistency of both sides of the BAE. For ideal A strings of length m the rapidities are A™J = A£ + (n + 1 - 2j)iu + 0(exp(-SN)). Here u = c/2, a enumerates the strings of the same length m, and j = 1 , . . . ,m counts the As involved in the ath A string of the length m, A™ is the real part of the string. Taking logarithm of the BAE (2) by using string hypothesis we arrive at the following discrete Bethe ansatz equations
1-KIJ = kjL + 2^2e2(fc,- -ki)-J2Q"(ki I
2nJ: = 2 £ On(A£ - *,) - 2 £ l
where 0n(,x) = tan
1
(x/nu)
~ A")
an
AnltQt(Xna - Xlb)
bl,tytO
and
1, fori = n + 1,17i — /| Anit = { 2, for i = n + 1 - 2, • • •, \n - l\ + 2 0, otherwise. 286
(3)
Thermodynamics of Two Component Bosons
2139
Ij and J™ play the role of quantum numbers for charge rapidity and isospin rapidity respectively. In order to guarantee linear independence of wave function, all quantum number within a given set of {7} as well as that in {J} should be different. An arbitrary quantum number may be either in the set or not in the set. The former is called a root, the later is called a hole. In thermodynamic limit, the distribution of charge rapidities becomes dense and it is useful to introduce the density function for charge roots and holes respectively. We denote with p(k) and ph(k) the density function of charge roots and holes, in a similar way, with
=
(l/L)dI{k)/dk (l/L)dJn(X)/d\.
(4)
Then from eqs. (3) we obtain a set of coupled integral equations. p + ph=
±+K2(k)*p(k)-J2Kn(k)*an(k) n
h
on + o n= Kn(\) * p(X) - J2 AnitKt{\)
* ai(X)
(5)
where Kn(x) — nu/ir(n2u2 + x2), and * denotes the integral convolution. In terms of the density functions of charge and isospin roots, the kinetic energy per length has the form Ef./L = j k2p(k)dk, the total number of down isospins is M/L = ^2nn Jan(X)dX and the particle density of the model is D = N/L = J p(k)dk. If we consider the energy arising from the external field 0 which is the Rabi field in two-component BEC experiments, the internal energy of the model is E/L = f{k2 - Q)p(k)dk + J ]
2n0
/ andX.
(6)
And with the help of the approach first introduced by Yang and Yang, 10 the entropy of the present model at finite temperature is S/L = f[(p + ph) ln(p + / ) - plnp -ph \nph}dk
(7)
+ J2 /[(CT« + *£) l n K + o£) - ov, In(7„ - ohn In <%]d\. n
J
The Gibbs free energy of the model then is defined by F = E — TS — pN, where p, is the chemical potential. In order to obtain the thermal equilibrium, we minimize the free energy with respect to all the density functions subjects to the constraint (5). In addition, the total number of particles, the magnetization are kept to constant. For this purpose, the chemical potential p and external field O play the role of Lagrange multipliers. It is useful to define the energy potential for charge sector and isospin sector: K(jfe) = e e(fc ) /T = Vn(X)
=ahn{X)/an{X). 287
ph{k)/p{k) (8)
2140
S.-J. Gu et al.
Applying the minimum condition SF = 0 gives rise to a revised version of GaudinTakahashi equations TlnAv -
e(k) = k2-fi-Q-
TK2{k) * ln[l + KT 1 ]
-Tj2Kn(k)*H^-+V~1} n
lnr/i = —sech(7rA/2u) * ln[(l +
K_1)(1
+ %)]
ln77„= —sech(7rA/2u)*ln[(l+»fc,_i)(l + »M+i)]. 4u And these equations are completed by the asymptotic conditions lim [lnr]n/n} = 2x
(9)
(10)
where x — fi/T. Eqs. (9) can be solved by iteration. Note that eqs. (5) together with eqs. (9) completely determine the densities of charge roots and isospin roots in the state of thermal equilibrium. The Helmholtz free energy F = E — TS is given by F=
f ln[l + e-elT\dk.
nN-—
(11)
The above approach called TBA is universal for discussing the thermodynamics of one dimensional integrable model. Once eqs. (9) are solved, all thermodynamic quantities can be obtained from eq. (11) in principle. 3. Magnetic Property: Zero and High Temperature Limit: The state at zero temperature is the ground state. The Fermi surface is determined by e{kp) = 0. Since there is no hole under Fermi surface, we can take the energy potential K = ph/p as zero. As a result, from eqs. (9), it is easy to see r]n —> oo, and M = 0, the "ferromagnetic" ground state. The first equation of eqs. (9) becomes e0(k) = k2 - ix - il + K2(k) * e0(k)
(12)
which gives the solution of dressed energy, and the ground-state energy may be given in terms of €o
1
rhF
E0/L = — / e0(k)dk. (13) 27r J-kF Consequently, the ground state of ID SU(2) bosons is an isospin "ferromagnetic" state, which coincides with the analysis of Li et al. 4 Then the property of the model at T = 0 is the same as that of scalar bosons in one dimension which has been discussed extensively by Lieb and Liniger. 6 In the isospin space, however, the SU(2) symmetry of whole system around the ground state is broken. 288
Thermodynamics
of Two Component Bosons
2141
In the high temperature limit T —> oo (free isospins), however we can assume that all functions r]n(\) are independent of A. Then eqs. (9) can be written as follows,
r)l = (1 + m) nl=
(l + Tfe-xXl + Tfc+i)
(14)
where we have neglected the term (1 + K - 1 ) in the second equation of eqs. (9). The solution of j]n are then constants fixed by the field boundary condition (10) to be Vn =
'sinh(n + l)x]2 sinhx
1
After perform the Fourier transformation on eqs. (5), we get the solution of the densities of A n-strings,
"=
^sech[7rA/2u]*[^1+«7ti].
(16)
If we assume that an andCT£are independent of A or let c = 0, the total number of down isospins has the form,
£nan = ^ - ^ . „
m
e - ^
(17)
n
where nm is maximal length of A strings. In the absence of Rabi field, we have M/N = 1/2, the system at high temperature is a quasi "paramagnetic" state. If the external field 0 is small, expanding eq. (17) for small field x and integrating the equation over A space, we get the magnetization of the model. Let Mm be the total number of isospin rapidities in all n m -strings,
where the first term in the parentheses arises from self-magnetization, while the others are contributed by Rabi field. Eq. (18) indicates that the magnetic property of the model in high temperature regime dominated by Curie's law x oc 1/T. 4. Strong and Weak Coupling Limit: When 77 —>• 00, Kn(k) = 0, from eqs. (9) we have e = k2 - il - /i-
(19)
The free energy of the system (11) at low temperature now can be solved by integration by part, F,L
2 1,3/2
= ,D--y-^
289
2^2
r v
+ WJ-2
(20)
2142
S.-J. Gu et cd.
where the external field is set to zero. We can not deduce the specific heat directly from the free energy obtained above because the chemical potential is a function of temperature. Prom eqs. (5), the density of charge rapidity has the form -
1
1
Clearly, at zero temperature, the Fermi surface is just the square root of the chemical potential, so we have ^o = 7r 2 D 2 . At low temperature, however, it is determined by D = N/L = / p(k)dk. After integration, we have a temperature dependent chemical potential 2^2-, - 2
1
(22)
24/x2. J
Then the free energy becomes 2
7T
2j>2
Since by thermodynamics S = —OF/dT and Cv — TdS/dT, heat at low temperature is Fermi-liquid like S = CV = ^ .
we find the specific
(24)
It is the same as the result of one-component case, since for the strong coupling limit the isospin and the charge are decoupled completely, the contribution of isospin to the free energy vanishes. In order to discuss the possibility of the existence of BEC, we consider the problem in weak coupling limit u —> 0. And isospin-isospin reaches its maximal correlation. At low temperature, however, we do not take string hypothesis for simplicity. Because \im.c^oKn(x) = 5(x), together with eqs. (5) and eqs. (9), we obtain PW
_ 1 (3e»-l)(e-+l) '27r(3e^o+l)(l_e-eo)
^5)
where so = (k2 — n)/T. The positive definition of p(k) requires that the chemical potential is negative. As we known the density of scalar boson is 2np = 1/(1 — e~ e °) which prevents the BEC in ID and 2D system because of the infrared divergence. However, the density function (25) still does not resolve this problem. Consequently, BEC does not happen in this model yet. 5. Conclusion and Acknowledgment To summarize, we discussed the general thermodynamics of one dimensional SU(2) bosons with J-function interaction by using the strategy of TBA. It was shown that the ground state is an isospin "ferromagnetic" state which differs from the ground state of ID fermions, while at high temperature, it is "paramagnetic" state and 290
Thermodynamics of Two Component Bosons 2143 t h e magnetic property is d o m i n a t e d b y Curie's law. In s t r o n g coupling limit, we obtain the exact expression of t h e d e p e n d e n c e of chemical potential, entropy and specific heat on temperature which a r e Fermi-liquid like, while in weak coupling limit, we found the infrared divergence of charge roots density function prevents t h e existence of B E C . This work is supported by T r a n s - C e n t u r y Training P r o g r a m Foundation for the Talents and EYF98 of China Ministry of E d u c a t i o n . S J G t h a n k s D.Yang for helpful discussions.
References 1. J. E. Williams, M. J. Holland, Nature 4 0 1 , 568(1999); J. E. Williams, M. J. Holland, C. E. Wieman, and E. A. Cornell, Phys. Rev. Lett. 83, 3358(1999). 2. T. L. Ho, Phys. Rev. Lett. 8 1 , 742(1998). 3. P. P. Kulish, Physica 3D 1, 246 (1981). 4. Y. Q. Li, S. J. Gu, Z. J. Ying and U. Eckern, submitted to P R B , [cond-mat/0107578]. 5. E. L. Lieb and F. Y. Wu, Phys. Rev. Lett 20, 1445 (1968). 6. E. H. Lieb and W. Liniger, Phys. Rev. 130, 1605 (1963). 7. L. D. Faddev, in Recent Advances in Field Theory and Statistical Mechanics, ed. J.-B. Zuber and R. Stora (Elsevier, Amsterdam 1984) p. 569. 8. C. N. Yang, Phys. Rev. Lett. 19. 1312 (1967). 9. E. H. Lieb, Phys. Rev. 130, 1616 (1963). 10. C. N. Yang and C. P. Yang, J. Math. Phys. 10 1115 (1969). 11. M. Takahashi, Prog. Theor. Phys. 4 7 , 69 (1972).
291
Internationa] Journal of Modern Physics B, Vol. 16, Nos. 14 & 15 (2002) 2145-2151 © World Scientific Publishing Company
R-MATRICES A N D T H E T E N S O R P R O D U C T G R A P H M E T H O D
MARK D. GOULD YAO-ZHONG ZHANG Centre for Mathematical Physics, Department of Mathematics, Brisbane, Qld 4072, Australia
University
of
Queensland,
Received 12 November 2001
A systematic method for constructing trigonometric R-matrices corresponding to the (multiplicity-free) tensor product of any two affinizable representations of a quantum algebra or superalgebra has been developed by the Brisbane group and its collaborators. This method has been referred to as the Tensor Product Graph Method. Here we describe applications of this method to untwisted and twisted quantum affine superalgebras.
1. Introduction The (graded) Yang-Baxter equation (YBE) plays a central role in the theory of (supersymmetric) quantum integrable systems. Solutions to the YBE are usually called R-matrices. The knowledge of R-matrices has many physical applications. In one-dimensional lattice models, R-matrices yield the Hamiltonians of quantum spin chains.1 In statistical mechanics, R-matrices define the Boltzmann weights of exactly soluble models2 and in integrable quantum field theory they give rise to exact factorizable scattering S-matrices.3 So the construction of R-matrices is fundamental in the study of integrable systems. Mathematical structures underlying the YBE and therefore R-matrices and integrable models are quantum affine (super)algebras. A systematic method for the construction of trigonometric R-matrices arising from untwisted and twisted quantum affine (super)algebras has been developed in Refs. 4-9 (see also Ref. 10 for rational cases). This method is called the Tensor Product Graph (TPG) method. The method enables one to construct spectral dependent R-matrices corresponding to the (multiplicity-free) tensor product of any two affinizable representations of a quantum algebra or superalgebra. In this contribution, we describe the TPG method in the context of untwisted and twisted quantum affine superalgebras. Quantum superalgebras are interesting since the tensor product decomposition often has indecomposables and integrable models associated with them may in some instances be interpreted as describing strongly correlated fermion systems. 11 ' 12 293
2146
M. D. Gould & Y.-Z.
Zhang
2. Quantum Affine Superalgebras and Jimbo Equation Let us first of all recall some facts about the affine superalgebra G^k\ k — 1,2. Let Go be the fixed point subalgebra under the diagram automorphism of Q of order k. In the case of k = 1, we have Go = G- For k = 2 we may decompose G as Go © Gi, where [£i] C G\- Let tp be the highest root of Go = G for k = 1 and 0 be the highest weight of the ^-representation G\ for A; = 2. Quantum affine superalgebras Uq[G^] are g-deformations of the universal enveloping algebras U[Q^] of 5 ^ • We shall not give the defining relations for UglG^}, but mention that the action of the coproduct on its generators {hi, e^, fi, 0 < i < r} is given by A(/ii) = hi® 1 + 1® hi, A(ei) = ei®q^
+ q~^ ® eu
A(/i) = / j
(1)
Define an automorphism Dz of Uq[G^] by £»z(e,) = zk5<°ei,
= z-kSi°fi,
Dz(fi)
Dz(hi) = K
(2)
Given any two minimal irreducible representations TT\ and 7r^ of Uq[Go] and their affinizations to irreducible representations of Uq[G^], we obtain a one-parameter family of representations A? of Uq[G^] on V(X) ® V(fj,) defined by A^(a)=7rx®7r^((I?z®l)A(a)),
Va € CM0(fc)],
(3)
x
where 2 is the spectral parameter. Let R ^(z) be the spectral dependent R-matrices associated with ir\ and 7rM, which satisfies the YBE. Moreover it obeys the intertwining properties:
i ? ^ ) A y a ) = (Ar)ya)i^(z) 13
(4)
XlJ,
which, according to Jimbo, uniquely determine R (z) up to a scalar function of z. We normalize Rxi*{z) such that Rx'1{z)R'iX{z-1) = I, where Rx»(z) = P Rxi*{z) with P : V(X) ® V(fj.) —> V(/J,) ® V(X) the usual graded permutation operator. In order for the equation (4) to hold for all a € Uq[G^] it is sufficient that it holds for all a G Uq(Lo) and in addition for the extra generator eo. The relation for eo reads explicitly Rx»(z)
(z7rx(eo)®Trli(qh°/2)+*x(q-h°/2)®^(eo))
= (^(e0)®irx(qh°'2)
+ znfi(q-h°/2)®7rx(e0))Rx»(z).
(5)
Eq.(5) is the Jimbo equation for Uq[Q^]. 3. Solutions to J i m b o Equation and Tensor Product Graph Method Let V(X) and V([i) denote any two minimal irreducible representations of Uq[G^\. Assume the tensor product module V(A) ® V(fi) is completely reducible into irre294
R-Matrices
and the Tensor Product Graph Method
2147
ducible Uq[Go\-mo&\Aes as V{\)®V{ii)
= ($V{v)
(6)
and there are no multiplicities in this decomposition. We denote by P^ the projection operator of V[X) ® V{y) onto V{v) and set P* M = R^(l) P^ = Ppx Rx»(l). We may thus write #"(*) = X > ( s ) P ^ ,
p„(l) = l.
(7)
Following our previous approach, 5 the coefficients pu(z) may be determined according to the recursion relation P {Z)
"
fWt
+
= zfM
e^zqW*
+ wfVV*""^
(8)
which holds for any v ^ v' for which P^
( ^ ( e o ) ® KMh°'2))
P?
± 0.
(9)
Here C(u) is the eigenvalue of the universal Casimir element of Go on V{v) and eu denotes the parity of V(u) C V"(A)
l±zqa •
0
,
(10)
so that the relation (8) may be expressed as
^./m^m)
MJ,
(11)
To graphically encode the recursion relations between different pu we introduce the Extended TPG for Uq[G(1)] and Extended T w i s t e d T P G for Uq[G{2)}. Definition 1: The Extended T P G associated to the tensor product V(X)
and
e(i/)e(«/) = - 1 .
(12)
The condition (12) is a necessary condition for (9) corresponding to Uq[G^] to hold. Definition 2: The Extended Twisted T P G which has the same set of nodes as the twisted TPG but has an edge between two vertices v ^ v' whenever V(i/) C V(6)
(13)
2148
M. D. Gould & Y.-Z.
Zhang
and _ J + 1 if V(v) and V(v') are in the same irreducible representation of Q \ — 1 if V(v) and V(i/) are in different irreducible representations of Q. (14) The conditions (13) and (14) are necessary conditions for (9) corresponding to Ug[gW] to hold. We will impose a relation (8) for every edge in the extended (twisted) TPG but we will be imposing too many relations in general. These relations may be inconsistent and we are therefore not guaranteed a solution. If however a solution to the recursion relations exists, then it must give the unique correct solution to the Jimbo's equation. 4. Examples of R-matrices for
Uq[gl(m\n)^]
Throughout we introduce {ej}^ =1 and {£j}? = i which satisfy (ej,ej) = 6ij, (6i,Sj) = —Sij and (u,Sj) = 0. As is well known, every irreducible representation of Uq[gl(m\n)] provides also an irreducible representation for Uq[gl(m\n)^]. Here, as examples, we will construct the R-matrices corresponding to the following tensor product: rank a antisymmetric tensor with rank b antisymmetric tensor of the same type. Without loss of generality, we assume m>a>b and the antisymmetric tensors to be contravariant. The tensor product decomposition is V(\a)®V(\b)
= ($V(Ac)
(15)
c
where, when a + b < m, b
a+c
Ab = ^ € i ,
b—c
Ac = ^ e i + ^ e i ,
i=l
i=l
c = 0,l,---,6
(16)
j=l
and when a + b > m, a+c
b—c
Ac=^ei + ^ei, t=i m
c = 0,1, • ••,m-
Ac = ^ C J + ^ e j + (a + c - m ) 5 i , t=i
a
»=i b—c
c — m-a
+ l,---,b
(17)
»=i
The corresponding TPG is V(\a)®V(\b)
(18) A0 Ai A 6 _i Afc which is consistent; such is always the case when a graph is a tree (i.e. contains no closed loops). From the graph we obtain b
#«•**(x) = ^2Y[{2i + a-b)_ pfcXh c=0 i = l 296
(19)
R-Matrices and the Tensor Product Graph Method 2149
The a = b = 1 case had been worked out before, which is known to give rise to the Perk-Schultz model R-matrices. 14,15 5. Examples of R-matrices for C/9[gZ(n|n)^2)] To begin with, we introduce the concept of minimal representations. By minimal irreducible representations of Q, we mean those irreducible representations which are also irreducible under the fixed subalgebra QQ. We can determine R-matrices for any tensor product V(\a)®V{\b) of two minimal representations V(Xa), V(Xb) of Uq[gl(m\n)^], where V(Xa) is also irreducible model under Uq[osp(m\n)} with the corresponding Uq[osp(m\n)\ highest weight Aa = (6|a, 6). Recall that for our case QQ = osp(m\n) and 6 = 8i + 82. Below we shall illustrate the method for the interesting case of a — b, m = n > 2, where an indecomposable appears in the tensor product decomposition. The decomposition of the tensor product of two minimal irreducible representations of Uq[osp(m\n)]:9 a
V(Xa)®V(Xb)
c
= Q)(£)V(k,a
+ b-2cy,
(20)
c=0 k=0
here and throughout V(a, b) denotes an irreducible Uq[osp(m\n)} module with highest weight Xat(, = (0|a + 6, a, 0). Note that one can only get an indecomposable in (20) when m = n > 2 and a + b - 2c = 0. Since a < b, c < a, this can only occur when a = b and c = a. In that case the C/q[osp(m|n)]-modules V(k,0), k — 0,1, will form an indecomposable. From now on we denote by V this indecomposable module, and write the Uq[osp(n\n)} module decomposition (20) as V{\a)®V{\a) = @V(v)®V,
(21)
where the sum on v is over the irreducible highest weights. Note that V contains a unique submodule V^#i + ^2) which is maximal, indecomposable and cyclically generated by a maximal vector of weight 81+82 such that V/V{8\+52) = V(0j0) (the trivial f/q[osp(n|n)]-module). Moreover V contains a unique irreducible submodule V(0|0) C V(8i +82)- The usual form of Schur's lemma applies to V{8\ +82) and so the space of Uq[osp(n\n)] invariants in End(V) has dimension 2. It is spanned by the identity operator I together with an invariant ./V (unique up to scalar multiples) satisfying NV = V{0\6)cV(81
+ 82),
NV(81 + 82) = (0).
(22)
It follows that TV is nilpotent, i.e. TV2 = 0. We can show9 that the minimal irreducible Uq[osp(n\n)} modules, V(Xa), with highest weight Aa, are affinizable to carry irreducible representations of Uq[gl(n\n)(2'] We now determine the extended twisted TPG for the decomposition given by (21). 297
2150 M. D. Gould & Y.-Z. Zhang
We note that V can only be connected to two nodes corresponding to highest weights (opposite parity), + 62) (same parity),
(c, k) = (a - 1,0) (c, k) — (a, 2).
,
We thus arrive at the extended twisted T P G for (21), given by Figure 1. k =0 c= 0
c=l
c= 2
k =a—1 c= a— 1 A; = a c= a Fig. 1. The extended twisted TPG for Ug[gl(n\n)^] (n > 2) for the tensor product V(Xa) ® y(A a ). The vertex labelled by the pair (c, k) corresponds to the irreducible Uq[osp(n\n)] module V(k,2a — 2c) except for the vertex corresponding to c = a, k = 1, which has been circled to indicate that it is an indecomposable [79[osp(n|n)]-module. It can be shown that the extended twisted TPG is consistent, i.e. that the recursion relations (8) give the same result independent of the path along which one recurs. To prove this it suffices to show for each closed loop of four vertices in the graph, that the difference in Casimir eigenvalues for osp{n\n) along one edge equals the difference along the opposite edge. Let Py = PyaXa be the projection operator from V(A n )
+ Pv{z)Pv 298
+ J2 P»(z)p»-
(24)
R-Matrices and the Tensor Product Graph Method
2151
T h e coefficients pv{z) can b e obtained recursively from the extended twisted T P G . However, t h e coefficients PN{Z) a n d pv{z) cannot be read off from t h e extended twisted T P G since the corresponding vertex refers to an indecomposable module. R a t h e r t h e y are determined by t h e a p p r o a c h 1 6 to Uq[gl(2\2)^]. The result is 9 a ' c ' c—fe
R{z) = pN{z)N
+ pv(z)Pv
+ E E I 1 ^ -
2a
>+
c=0 fe=0 j = l c
Yl(i-2a-l)-P{2a-2c+k)s1+kS2,
(25)
i=l
where t h e p r i m e s in the sums signify t h a t t e r m s corresponding to c = a w i t h k = 0,1 are o m i t t e d from t h e sums, a n d pv(z),
PN(Z) are given by
2 a—1 q
PN{z)
= (-l)V
a—1
j=i a 2
t=i
\ ^ z Pv{z). 1+2
(26)
Acknowledgements This work h a s b e e n financially s u p p o r t e d by the Australian Research Council.
References 1. E. K. Sklyanin, L. A. Takhtadzhyan and L. D. Faddeev, Theor. Math. Phys. 40, 194 (1979). 2. R. J. Baxter, Exactly Solved Models in Statistical Mechanics, Academic Press, London (1982). 3. A. B Zamolodchikov, Al. B. Zamolodchikov, Ann. Phys. 120 (1979) 253. 4. R. B. Zhang, M. D. Gould and A. J. Bracken, Nucl. Phys. B354, 625 (1991). 5. G. W. Delius, M. D. Gould and Y.-Z. Zhang, Nucl. Phys. B432, 377 (1994). 6. G. W. Delius, M. D. Gould, J. R. Links and Y.-Z. Zhang, Int. J. Mod. Phys. A10, 3259 (1995). 7. G. W. Delius, M. D. Gould and Y.-Z. Zhang, Int. J. Mod. Phys. A l l , 3415 (1996). 8. G.M. Gandenberger, N. J. MacKay and G. M. T. Watts, Nucl. Phys. B 4 6 5 , 329 (1996). 9. M. D. Gould and Y.-Z. Zhang, Nucl. Phys. B566, 529 (2000). 10. N. J. MacKay, J. Phys. A 2 4 , 4017 (1991). 11. F. H. L. Essler, V. E. Korepin and K. Schoutens, Phys. Rev. Lett. 68, 2960 (1992). 12. A. J. Bracken, M. D. Gould, J. R. Links and Y.-Z. Zhang, Phys. Rev. Lett. 74, 2768 (1995). 13. M. Jimbo, Int. J. Mod. Phys. A 4 , 3759 (1989). 14. J. H. H. Perk and C. L. Schultz, Phys. Lett. A84, 407 (1981). 15. V. V. Bazhanov and A. G. Shadrikov, Theor. Math. Phys. 73, 1302 (1988). 16. M. D. Gould, J. R. Links, I. Tsohantjis and Y.-Z. Zhang, J. Phys. A30, 4313 (1997).
299
International Journal of Modern Physics B, Vol. 16, Nos. 14 & 15 (2002) 2153-2159 © World Scientific Publishing Company
FREE FIELD A N D P A R A F E R M I O N I C REALIZATIONS OF TWISTED su(3)£. 2) C U R R E N T A L G E B R A
XIANG-MAO DING * M A R K D. GOULD YAO-ZHONG ZHANG Centre for Mathematical Physics, Department of Mathematics, Brisbane, Qld 4072, Australia
University
of
Queensland,
Received 13 November 2001 Free field and twisted parafermionic representations of twisted su(3), current algebra are obtained. The corresponding twisted Sugawara energy-momentum tensor is given in terms of three (/3,7) pairs and two scalar fields and also in terms of twisted parafermionic currents and one scalar field. Two screening currents of the first kind are presented in terms of the free fields.
1. Introduction Infinite dimensional algebras, such as Virasoro algebra and affine algebras are algebraic structures in conformal field theories (CFT) in two dimensional spacetime. 1-3 They also play a central role in the study of string theory. 4 It is well-known the untwisted current algebra can be realized by at least two different ways: one is the free field representation, 5 and the other is the parafermion representation. 6 The free field realization is a common approach used in conformal field theories. 5 The free field representations for untwisted affine algebra have been extensively studied. The simplest untwisted case su(2)^ was first treated in Ref. 5, and the generalization to su(n)^ was given in Refs. 7-17. The Zk parafermions are generalizations of the Majorana fermions, and the Z\. parafermion models are extensions of the Ising model, which corresponds to the case fc = 2. 1 9 , 2 0 Parafermions are related to the exclusion statistics introduced by Haldane. 21 The recently study shows that twisted affine algebras are useful in the description of the entropy of Adsz black hole. 18 However, little is known for free field and parafermionic representations of twisted affine algebras. So it is an interesting problem to investigate such realizations. In this contribution we consider the simplest twisted affine algebra sw(3)^.'. We construct two different realizations for this algebra: one is free field representation, 'Permanent address: Institute of Applied Mathematics, Academy of Mathematics and System Sciences; Chinese Academy of Sciences, P.O.Box 2734, 100080, China. 301
2154
X.-M. Ding, M. D. Gould & Y.-Z. Zhang
which is another version of one given in Ref. 22, and another is twisted parafermionic representation by using the twisted parafermionic currents proposed in Ref. 23. Moreover we give the screening currents in terms of the free fields. 2. Twisted current sw(3)^ ' algebra We consider the simplest twisted affine Lie algebra
(2)
SM(3))J.
. We decompose su(3)
as su(3) = go © ffi
(1)
where 50 = su(2) is the fixed point subalgebra under the automorphism and #1 is the five dimensional representation oi go; go and g\ satisfy [gi,gj] C g(i+j) mod 2- Using the notation in Ref. 22, we chose e, / and h to be the bases in go and e, / , E, F and h the bases of g\. Then the commutators of su(3)k can be expressed as >X,zn®Y]
= z m+n
,
(X\Y) [X,Y] + 2kmSm+n,o-
(2)
Where m £ Z if X € go, and m € Z + \ if X € gi. Denote the currents corresponding to e, h, f by j+(z), j°(z), j~(z), and to e, h, / , respectively. Then (2) can be written in terms of the following OPE's: j+(z)j
Ak (z — w)2
(w)
+
1 r j » + -., (z — w)"
Ak
J+(z)J-(w) J++(z)J-(w)
J°(z)J±(w)
1
=
^ + T-±-J°(w) + ..., (z — w)z (z — w) = T-^L- 2 + j-^—fiw) + ..., (z — w) (z — w) 8k ±2 • j ± H + ..., j°(z)j°(w) (z-w] (z — w)2 ±6 ±6 ± = J±(w) j H + ... (z — w) (z — w)
J+(z)J—(w)
= 7-l—j-(w) (z-w)' =
j+{z)J-(v,)
= j ^ J ° W
j+{z)J—{w)
= j-=*—J-(w) +. (z — w) ±2 = J±(w) + ... (z — w) 2Ak = r + .... (z — w)
fiz^iw) J°{z)J°{w)
(z — w)
J++(w)
J-{z)J++{
+.
j+{z)J+{w)
j+(w)+ ..... (z — w) -2 j (z)J (w) = J—{w) + ..., (z — w)
+. + ...
+ ...,
J+(z)j-(w)
=
j-(z)J++{w) fiz^iw)
302
(z — w) =
=
J°(w) +. — J+(w)
±4 j (z — w)
± ±
H + ...:
Free Field and Parafermionic
Realizations
2155
All other OPE's contain trivial regular terms only. Here and throughout "..." stands for regular terms. 3. Wakimoto free field realization of the twisted affine currents To obtain a free field realization of the twisted su(3yk' currents, we follow the procedure adopted in Ref. 22 but begin with a different Fock space. The Fock space is constructed by the repeated actions of / , / , F on the highest weight state v\ with highest weight A. The highest weight states are determined by evA = evA = EvA = 0,
hvA = (A, ai)v A ,
hvA = (A, a2)vA
(3)
where a\ and a2 are roots associated with h and h. Set \n, m, I > = fnFmflVAThis choice is different from the one used in Ref. 22. As we shall see, this choice gives another free field realization of the twisted current algebra. Introduce three /3j pairs and two scalar fields (j>a, a = 1, 2. (/%; ji) pairs have conformal dimension (1; 0). ft(2)7; H = - T i W f t H = - - ^ - , *, j = 0,1,2 z—w 4>a{z)4>b{w) = -25a,b ln(z - w), a, b = 0,1
(4)
Introduce the notation ei = | ( 1 , 1 ) ; e2 = - ^ ( 1 , - 1 ) ; and l> = (4>o,4>±). Then we have e\ • $(z)ei •
e2 • §>{z)e2 • $(w) = - 3 ln(z — w); e*i • $(z)e2 • $>(w) = 0.
We find the Wakimoto free field realization of su(3)jj.' in terms of the eight free fields: j+(z)=
A>(z)+2ft(z)7i(*),
j°(z) = 2p0(z)l0(z) j~(z)
+ 2(31{z)ll{z)+Ap2{z)l2{z)
+ — (ei • a+
id$(z)),
= -Po(z) (7o2W + 3 7 l 2 W) - 2fa{z) (370(2)71 W - 72(2)) -2/32(z) (37g(2)7i W + 7?(*)) - 4(fc + l)57o(«) (ei • id$(z))nf0(z)
j\z)
(e2 • id$(z))7i(;z),
= 6/3o(2)7iW + 6/?i(2)70(2) + 6ft(z) (7o2(2) + 7i2(2))
+ — (e2-id$(z)); a+ J+(Z) = /3l(2),
(5)
J - ( z ) = -2A)(z) (70(2)71(2) +72(2)) - A(2) (37o2(2) +7i 2 (2)) -4ft(z) (7o3(2) +71(2)72(2)) -4fc9 7 i(2), (ei • ia$(z))7!(2;) 303
(e2 •
id$(z))j0(z);
2156 X.-M. Ding, M. D. Gould & Y.-Z. Zhang
J—(z)
= 20o(z) ( 7 g(*)7i(*) +7i 3 (z) -
27o(zh2(z))
2
- 2 f t (z) (
= T^zftz-"-1;
Ja(z) = ZnzZ+1/2JZz-n-\
(6)
4. Twisted parafermions and parafermionic realization In this section we use the twisted parafermionic currents proposed in Ref. 23 to give another realization of the twisted sw(3)J.' currents. First, we recall the twisted parafermion current algebra given in Ref. 23 reads,
(z — wy
z—w
_„A"72fc ==fJML Wt>i.H(z-")""* r wz ^ + P H where I, I' — ± 1 and I, I' — 6, ± 1 , ±2; sij', ej p and et p are structure constants. If we denote ipi or ip^ by V&a, then we can rewrite the above relations as: 00
- w)ab'2k
*a(z)Vb(w)(z
=
£ (* - ™)n[*a^\-n, n=-2
(8)
For consistency, ea,b must have the properties: ea,b = Sb,a = —£-a,-& = £-a,a+6 and ea ~a = 0. As a CFT, we can set the structure constants as e
l,-2
— e
l,l —e-l,2
_
1
—
\
rri
3 -1,-1 - c o , I - - y ^fe-
(9)
Then we may obtain a representation of the twisted su(3)k current algebra with the help of the twisted parafermionic currents. The result is j+(z) = 2Vktl>i(z)e^Mz),
J~(z) = 2 v / f c ^ _ 1 ( z ) e - ^ ' W z ) , 304
Free Field and Parafermionic Realizations 2157 f(z)
= 2V2Jfet0fo(z), = 2Vki>_i(z)e-^Mz\
J-(z) J—(z)
2Vk^i(z)e^Mz),
J+(z) = J++(z)
= 2y/kip_i(z)e-i^^z\
2Vk^(z)ei^Mz),
=
J°(z) = 2v/6fcVo(2)-
where
5. Twisted stress energy tensor It is well known that Virasoro algebras are related to currents algebras via the so called Sugawara construction. In the present case, the twisted Sugawara construction of the energy-momentum tensor is given by T(z)
X
\f{z)f{z)
8(fc + 3)
+ \j°{z)J\z)
+2J-(z)J+(z)
+
2j-(z)j+(z)
+ 2J—(z)J++(z)}
:,
(10)
where : : implies the normal ordering. The above expression can be rephrased through the ffy pairs and the scalar field 3>. We obtain T(z) = - : \p0(z)d7o(z)
+ (31(z)d11(z)
+ - : (Si • id$(z))
+ f32(z)dl2(z)j
: + - : (e 2 • id$(z))
:
:
- 4 a + (gi • id2$(z)V
(11)
On the other hand, the energy-momentum tensor in the twisted parafermionic realization is given by T(z) = T+ -
: d
(12)
where 2fc + 6
( 13 )
E[*-*-Jo
is the energy-momentum tensor of the twisted parafermion currents obtained in Ref. 23. Following the standard practice, we get the OPE of the energy-momentum tensor,
n,m«) =
^
+
*n=L
{z — wp
+
[z — wY
«W + ..., z —w
where c = 8k/(k + 3) is the central charge for the Virasoro algebra. 305
(14)
2158
X.-M.
Ding, M. D. Gould & Y.-Z.
Zhang
6. T w i s t e d screening currents An important object in the free field approach is the screening current. Screening currents are a primary fields with conformal dimension 1, and their integration give the screening charges. They commute with the affine currents up to total derivatives. These properties ensure that screening charges may be inserted into correlators while the conformal or affine Ward identities remain intact. For the present case, we find the following screening currents S±(z) = : [2fo(z)7o(z)+Pi(z)±l3o(z)]S±(z)
:,
(15)
where S±(z) = e - ^ + ^ i - ^ W i ^ W ) .
(16)
The O P E of the twisted screening currents with the twisted affine currents are T(z)S±(w)
= dw
(-^—S±(w)
j+(z)S±(w)
=
...,
j°(z)S±{w)
=
...,
j-(z)S±(w)
= dw (±^L--L-S±(w)^
J++(z)S±(w)
=
...,
J (z)S±(w)
=
...,
J°(z)S±(w)
=
...,
J-(z)S±(w)
= dw
+
J—(z)S±(w)
^ _ 1 _ 5
= dw (4"—^\ Oil Z
±
H )
+ ...,
+...,
(17)
[ 7 o H T 7i N ] S±(w)) IV
+...; J
The screening currents obtained here are the twisted versions of the first kind screening currents. 7 Acknowledgements This work is financially supported by Australian Research Council. One of the authors (Ding) is also supported partly by Natural Science Foundation of China, and a Foundation from AMSS. References 1. A. A. Belavin. A. M. Polyakov, A. B. Zamolodchikov, Nucl. Phys. B241, 333 (1984). 2. V. G. Kac, Infinite Dimensional Lie Algebras, third ed., (Cambridge University press, Cambridge 1990). 3. Ph. Di Francesco, P. Mathieu and D. Senechal, Conformal Field Theory, (Springer, 1997). 306
Free Field and Parafermionic Realizations 2159 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20.
21. 22. 23.
J. Polchinski, String Theory, (Cambridge University Press, Cambridge 1998). M. Wakimoto, Commun. Math. Phys. 104, 605 (1986). D. Nemeschansky, Phys. Lett. B 224, 121 (1989). M. Bershsdsky and H. Ooguri, Commun. Math. Phys. 126, 49 (1989). B. Feigin and E. Frenkel, Russ. Math. Surv. 4 3 , 221 (1988). B. Feigin and E. Frenkel, Lett. Math. Phys. 19, 307 (1990). B. Feigin and E. Frenkel, Commun. Math. Phys. 128, 161 (1990). P. Bouwknegt, J. McCarthy and K. Pilch, Phys. Lett. B234, 297 (1990). P. Bouwknegt, J. McCarthy and K. Pilch, Commun. Math. Phys. 131, 125 (1990). P. Bouwknegt, J. McCarthy and K. Pilch, Prog. Phys. Suppl. 102, 67 (1990). K. Ito, Phys. Lett. B252, 69 (1990). G. Kuroki, Commun. Math. Phys. 142, 511 (1991). E. Frenkel, Free field realizations in representation theory and conformal field theory, hep-th/9408109. J. L. Petersen, J. Rasmussen and M. Yu, Nucl. Phys. B502, 649 (1997). S. Fernando and F. Mansouri, Phys. Lett. B 5 0 5 , 206 (2001). A. B. Zamolodchikov and V. A. Fateev, Sov. Phys, JETP 62, 215 (1985). J. Lepowsky and R. Wilson, Comm. Math. Phys. 62, 43 (1978); Invent. Math. 77, 199 (1984); Invent. Math. 79, 417 (1985); in: Vertex Operators in Mathematics and Physics, Proc. 1983 M.S.R.I. conference, ed. by J. Lepowsky et. al, (Springer-Verlag, New York, 1985), P.97. F. D. M. Haldane, Phys. Rev. Lett. 67, 937 (1991). X. M. Ding, M. D. Gould and Y. Z. Zhang, Twisted s/(3)jj.2) current algebra: Free field representation and screening currents, hep-th/0109009. X. M. Ding, M. D. Gould and Y. Z. Zhang, Twisted parafermions, hep-th/0110165.
307